1,812 454 5MB
Pages 295 Page size 235 x 382 pts Year 2003
This page intentionally left blank
Practical Statistics for Astronomers
Astronomy, like any experimental subject, needs statistical methods to interpret data reliably. This practical handbook presents the most relevant statistical and probabilistic machinery for use in observational astronomy. Classical parametric and non-parametric methods are covered, but there is a strong emphasis on Bayesian solutions and the importance of probability in experimental inference. Chapters cover basic probability, correlation analysis, hypothesis testing, Bayesian modelling, time series analysis, luminosity functions and clustering. The book avoids the technical language of statistics in favour of demonstrating astronomical relevance and applicability. It contains many worked examples and problems that make use of databases which are available on the Web. It is suitable for self-study at advanced undergraduate or graduate level, as a reference for professional astronomers, and as a textbook basis for courses in statistical methods in astronomy. jas per wa ll was, to 2003, Visiting Professor of Astrophysics and Director of Graduate Studies in the Department of Astrophysics at the University of Oxford. He obtained his Ph.D. in Astronomy at the Australian National University, Canberra, and has since been Head of Astrophysics at the Royal Greenwich Observatory, Director of the Isaac Newton Group of Telescopes, La Palma, and Director of the Royal Greenwich Observatory. Professor Wall has edited three books and published over 150 scientific articles on extragalactic radio sources, space distribution and cosmology, astronomy instrumentation, and statistics in astronomy. cha r les jen k in s has worked at Schlumberger’s Cambridge research lab since 1997, where he is a Principal Scientist working on innovations in oilfield telemetry and robotics. He obtained his Ph.D. in the Radio Astronomy group at the Cavendish Laboratory, Cambridge, and was an Astronomer at the Royal Greenwich Observatory for 14 years. Dr Jenkins has been involved in the commissioning of the Isaac Newton and William Herschel Telescopes as Project Scientist for numerous instruments, and latterly headed the New Projects Group and was Project Scientist for the tracking systems of the Gemini 8-m telescopes. His main research interests in astronomy were galaxy dynamics and adaptive optics.
PRACTICAL STATISTICS FOR ASTRONOMERS J. V. WALL UNIVERSITY OF OXFORD
C. R. JENKINS SCHLUMBERGER CAMBRIDGE RESEARCH
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo Cambridge University Press The Edinburgh Building, Cambridge CB2 8RU, UK Published in the United States of America by Cambridge University Press, New York www.cambridge.org Information on this title: www.cambridge.org/9780521454162 © J. V. Wall and C. R. Jenkins 2003 This publication is in copyright. Subject to statutory exception and to the provision of relevant collective licensing agreements, no reproduction of any part may take place without the written permission of Cambridge University Press. First published in print format 2003 eBook (EBL) ISBN-13 978-0-511-33697-3 ISBN-10 0-511-33697-7 eBook (EBL) ISBN-13 ISBN-10
hardback 978-0-521-45416-2 hardback 0-521-45416-6
ISBN-13 ISBN-10
paperback 978-0-521-45616-6 paperback 0-521-45616-9
Cambridge University Press has no responsibility for the persistence or accuracy of urls for external or third-party internet websites referred to in this publication, and does not guarantee that any content on such websites is, or will remain, accurate or appropriate.
In affectionate memory of Peter Scheuer (1930–2001) mentor and friend ‘2 + 2 5’
Contents
Preface Note on notation 1 1.1 1.2 1.3 1.4 1.5 2 2.1 2.2 2.3 2.4 2.5 3 3.1 3.2 3.3 3.4 3.5 4 4.1 4.2
page xi xiii
Decision How is science done? Probability; probability distributions Probability and statistics in inference Non-parametric or distribution-free statistical inference How to use this book Exercises Probability What is probability? Conditionality and independence . . . and Bayes’ theorem Probability distributions Inferences with probability Exercises Statistics and expectations Statistics What should we expect of our statistics? Simple error analysis Some statistics, and their distributions Uses of statistics Exercises Correlation and association The fishing trip Testing for correlation vii
1 3 4 6 7 8 9 10 11 15 17 24 32 34 37 37 41 43 49 51 52 54 54 57
viii
Contents
4.3 Partial correlation 4.4 But what next? 4.5 Principal component analysis Exercises 5 Hypothesis testing 5.1 Methodology of classical hypothesis testing 5.2 Parametric tests: means and variances, t and F tests 5.3 Non-parametric tests: single samples 5.4 Non-parametric tests: two independent samples 5.5 Summary, one- and two-sample non-parametric tests Exercises 6 Data modelling; parameter estimation 6.1 The maximum-likelihood method 6.2 The method of least squares: regression analysis 6.3 Bayesian likelihood analysis 6.4 The minimum chi-square method 6.5 Monte Carlo modelling 6.6 Bootstrap and jackknife 6.7 Models of models, and the combination of datasets Exercises 7 Detection and surveys 7.1 Detection 7.2 Catalogues and selection effects 7.3 Luminosity (and other) functions 7.4 Tests on luminosity functions 7.5 Survival analysis; censored data 7.6 The confusion limit Exercises 8 Sequential data – 1D statistics 8.1 Data transformations, the Karhunen–Loeve transform, and others 8.2 Fourier analysis 8.3 Statistical properties of Fourier transforms 8.4 Filtering 8.5 Correlating 8.6 Unevenly sampled data 8.7 Wavelets 8.8 Detection difficulties: 1/f noise Exercises
66 67 69 74 76 77 79 86 92 98 103 105 107 113 118 123 126 130 133 139 142 143 148 153 158 162 175 179 181 182 185 188 192 199 203 206 209 211
Contents
ix
9 9.1 9.2 9.3 9.4 9.5 9.6 9.7
Surface distribution – 2D statistics Statistics on a spherical surface Sky representation: projection and contouring The sky distribution of galaxies Two-point angular correlation function w(θ) Counts in cells The angular power spectrum Galaxy distribution statistics: interpretation Exercises Appendix 1 The literature Appendix 2 Statistical tables
214 214 219 220 221 229 236 243 244 246 250
References Index
265 271
Preface
Peter Scheuer started this. In 1977 he walked into JVW’s office in the Cavendish Lab and quietly asked for advice on what further material should be taught to the new intake of Radio Astronomy graduate students (that year including the hapless CRJ). JVW, wrestling with simple chi-square testing at the time, blurted out ‘They know nothing about practical statistics’. Peter left thoughtfully. A day later he returned. ‘Good news! The Management Board has decided that the students are going to have a course on practical statistics.’ Can I sit in, JVW asked innocently. ‘Better news! The Management Board has decided that you’re going to teach it’. So, for us, began the notion of practical statistics. A subject that began with gambling is not an arcane academic pursuit, but it is certainly subtle as well. It is fitting that Peter Scheuer was involved at the beginning of this (lengthy) project; his style of science exemplified both subtlety and pragmatism. We hope that we can convey something of both. If an echo of Peter’s booming laugh is sometimes heard in these pages, it is because we both learned from him that a useful answer is often much easier – and certainly much more entertaining – than you at first think. After the initial course, the material for this book grew out of various further courses, journal articles, and the abundant personal experience that results from understanding just a little of any field of knowledge that counts Gauss and Laplace amongst its originators. More recently, the invigorating polemics of Jeffreys and Jaynes, authors of standard works on probability, have been a great stimulus; although we have tried in this book not to engage too much with ‘old, unhappy, far-off things / and battles long ago.’ xi
xii
Preface
Amongst today’s practioners of practical statistics, we have had valued discussions with Mark Birkinshaw, Phil Charles, Eric Feigelson, Pedro Ferreira, Paul Francis, Dave Jauncey, Ofer Lahav, Steve Gull, Tony Lynas-Gray, Donald Lynden-Bell, Robert Laing, Louis Lyons, Andrew Murray, John Peacock, Chris Pritchett, Prasenjit Saha and Adrian Webster. We are very grateful to Chris Blake, whose excellent D.Phil. thesis laid out clearly the interrelation of 2D descriptive statistics; and who has allowed us to borrow extensively from this opus. CRJ particularly acknowledges the Bayesian convictions of the Real Time Decisions group at Schlumberger; Dave Hargreaves, Iain Tuddenham and Tim Jervis. Try betting lives on your interpretation of the Kolmogorov axioms. JVW is indebted to the Astrophysics Department of the University of Oxford for the enjoyable environment in which much of this was pulled together. The hospitality of the Department Heads – Phil Charles and then Joe Silk – is greatly appreciated; the stimulation, kindness, technical support and advice of colleagues there has been invaluable. Jenny Wall gave total support and encouragement throughout; the writing benefited greatly from the warmth and happiness of her companionship. CRJ wishes to acknowledge the support of Schlumberger Cambridge Research for the writing of this book, as part of its ‘Personal Research Time’ initiative. The encouragement of the lab’s director, Mike Sheppard, catalysed its completion. Program manager Ashley Johnson created the necessary space in a busy research group. Fiona Hall listened, helped with laughter and wise words through the long period of gestation, and took time out from many pressing matters to support that final burst of writing.
Notation
Here are some of the symbols used in the mathematical parts of this book. The list is not complete, but does include notation of more than localized interest. Some symbols are used with different meanings in different parts of the book, but in context there should be no possibility of confusion. alm : C:
the coefficients of a spherical harmonic expansion. usually the covariance (or error) matrix, characterizing a multivariate Gaussian. cl : the coefficients of the angular power spectrum. cov[x, y]: the covariance of two random variables x and y. D: the Kolmogorov-Smirnov test statistic. E[X]: the expectation or ensemble average. Also denoted . f, F : Probability density distributions and cumulative probability density distributions, respectively; in Chapter 8, Fourier pairs. F: a variable distributed according to the F distribution. H: the Hessian matrix. H0 , H1 : the null hypothesis and alternative hypothesis. K: the Kaplan-Meier estimator. L: intrinsic luminosity. L: the likelihood. N (S): the flux density distribution, or source count. N, n: usually the number of data. P (N ): the counts-in-cells probability of finding N objects in a cell. Pl : the Legendre polynomials.
xiii
xiv prob(. . .):
Note on notation
the probability of the indicated event. In the case of a continuous variable, the probability density. prob(A | B): the probability of A, given B. R: distance. r: the product-moment coefficient. R: the Rayleigh test statistic. S: the mean square deviation of a set of data; in Chapter 7, flux density. S: the test statistic for a particular orientation of the principal axis of the orientation matrix. Se : the sample cumulative distribution, as used in the Kolmogorov–Smirnov test. t: a variable distributed according to the t distribution. U: the Wilcoxon–Mann–Whitney test statistic. V, Vmax : the volume contained within R; the maximum volume, corresponding to the greatest distance consistent with an object still appearing in a catalogue. var[x]: the variance of a random variable x. w(θ): the two-point angular correlation function. X: the sample average of a set of data. X1 , X2 , . . .: usually a specific set of data; instances of possible data, denoted x. We try to keep to this distinction by using upper case for particular values and lower case for algebraic variables (although not with Greek letters, or statistics like t where lower-case is standard). y, z: the excess variance and skewness of clustered counts-in-cells. Ylm : the spherical harmonics. α : a vector, usually a vector of parameters. Γ: the Gehan test statistic. η: the luminosity distribution. κ: the Kendall test statistic. µ, σ: usually the mean and standard deviation of a Gaussian distribution; µ may also be the parameter of a Poisson distribution. µn : the nth central moment of a distribution.
Note on notation ρ:
the covariance coefficient of a bivariate Gaussian; in Chapter 7, the luminosity function. ς(θ, φ): the surface density of objects on the sky. σs : the sample standard deviation. φ: the space distribution. χ2 : a variable distributed according to the chi-square distribution.
xv
1 Decision
If your experiment needs statistics, you ought to have done a better experiment. (Ernest Rutherford)
Science is about decision. Building instruments, collecting data, reducing data, compiling catalogues, classifying, doing theory – all of these are tools, techniques or aspects which are necessary. But we are not doing science unless we are deciding something; only decision counts. Is this hypothesis or theory correct? If not, why not? Are these data self-consistent or consistent with other data? Adequate to answer the question posed? What further experiments do they suggest? We decide by comparing. We compare by describing properties of an object or sample, because lists of numbers or images do not present us with immediate results enabling us to decide anything. Is the faint smudge on an image a star or a galaxy? We characterize its shape, crudely perhaps, by a property, say the full-width half-maximum, the FWHM, which we compare with the FWHM of the point-spread function. We have represented a dataset, the image of the object, by a statistic, and in so doing we reach a decision. Statistics are there for decision and because we know a background against which to take a decision. To this end, every measurement we make, and every parameter or value we derive, requires an error estimate, a measure of range (expressed in terms of probability) that encompasses our belief of the true value of the parameter. We are taught this by our masters in the course of interminable undergrad lab experiments. Why? It is because no measured quantity or property is of the slightest use in
1
2
Decision
decision and therefore in science unless it has a ‘range quantity’ attached to it. A statistic is a quantity that summarizes data; it is the ultimate data reduction. It is a property of the data and nothing else. It may be a number, a mean for example, but it doesn’t have to be. It is a basis for using the data or experimental result to make a decision. We need to know how to treat data with a view to decision, to obtain the right statistics to use in drawing statistical inference. (It is the latter which is the branch of science; at times the term statistics is loosely used to describe both the descriptive values and the science.) Rutherford’s message appears uncompromisingly clear, but it can only hold in some specialized circumstances. For a start, astronomers are not always free to do better experiments. The laboratory is the big stage; the Universe is an experiment we cannot rerun. Attempting to understand astrophysics and cosmology from one freeze-frame in the spacetime continuum requires some reconsideration of the classical scientific method. This scientific method of repetition of experimentally reproduced results does not apply. We cannot reroll the dice, and anyway, repetition implies similar conditions. We are never at the same coordinates in spacetime. There is thus need for a certain rigour in our methodology. The inability to reroll dice has led and still leads astronomers into some of the greatest errors of inference. It becomes tempting to the point of irresistibility to use the data on which a hypothesis was proposed to verify that hypothesis.
The Black Cloud (Hoyle 1958). The Black Cloud appears to be heading for the Earth. The scientific team suggests that this proves the cloud has intelligence. Not so, says the dissenting team member. Why? A golf ball lands on a golf course which contains 107 blades of grass; it stops on one blade; the chances are 1 in 107 of this event occurring by chance. This is not so amazing – the ball had to land somewhere. It would only be amazing if the experiment were repeated to test the newly formulated hypothesis (e.g. the blade being of special attractive character; the golfer of unusual skill) so that the event were repeated. However, the importance of deciding if the Black Cloud knew about the Earth cannot await the next event or the sequence of events, and tempts the rush to judgement in which initial data, hypothesis and test data are combined; so in many instances in astronomy and cosmology.
EXAMPLE
1.1 How is science done?
3
A second difference for astronomers stems from the first – the remoteness of our objects and the inability to ‘rerun our experiments’ means that we do not necessarily know the underlying distributions of the variables measured. The essence of classical statistical analysis is (i) the formulation of a hypothesis, (ii) the gathering of hypothesis-test data via experiment, and (iii) the construction of a test statistic. But making a decision on the basis of the test statistic may demand that the sampling distribution or expectation of the statistic must be known before a decision can be made. To calculate this it is frequently essential to know the frequency distribution of the test statistic; how else could we decide if the value we got was normal or abnormal? It may well be the case that no one, physicist, sociologist, botanist, ever does know these underlying distributions exactly; but astronomers are worse off than most because of our necessarily small samples and our inability to control experiments, leading to poor definitions of the underlying distributions. It is thus the case that astronomers cannot avoid statistics and there are the following reasons at least for this unfortunate situation. (i) Error (range) assignment – ours, and the errors assigned by others: what do they mean? (ii) How can data be used best? Or at all? (iii) Correlation, hypothesis-testing, model fitting; how do we proceed? (iv) Incomplete samples, samples from an experiment which cannot be rerun, upper limits; how can we use these to best advantage? (v) Others describe their data and conclusions in statistical terms. We need some self-defence. (vi) But above all, we must decide. The decision process cannot be done without some methodology, no matter how good the experiment. Rutherford may not have known when he was using statistics. This is not a book about statistics, the values or the science. It is about how to get results in astronomy, using statistics, data analysis and statistical inference. Consider first how we do science in order to see at what point ‘statistics’ enter(s) the process. 1.1 How is science done? In simplest terms, each experiment goes round a loop which can be characterized by five stages:
4
Decision (1) Observe: record the data, or obtain the data. (2) Reduce: clean up the data to remove experimental effects, i.e. flat-field it, calibrate it. (3) Analyse: obtain the numbers from the clean data – intensities, positions. Produce from these summary descriptors of the data which enable comparison or modelling – descriptors which lead to reaching the decision which governed the design of the experiment; and which are statistics. (4) Conclude: carry through a process to reach a decision. Test the hypothesis; correlate; model, etc. (5) Reflect: what has been learnt? Is the decision plausible? Is it unexpected? At which experimental stage must re-entry be made to check? What is required to confirm this unexpected result? Or, what was inadequate in the experimental design? How should the next version be defined? Is an extended or new hypothesis suggested? Back to point (1).
This process is a loop and ‘experiments’ may begin at different points. For instance, we disbelieve someone else’s conclusions based on their published dataset. We enter at point (3) or even (4); and we may then go around the data-gathering cycle ourselves as a result. Or we look at an old result in the light of new and complementary ones from other fields – and enter at (5). All too often we use (3) to set up the tests at (4). This carries the charge of mingling hypothesis and data, as in the Black Cloud example. Table 1.1 summarizes the process. Points in Table 1.1 at which recourse to statistics or to statistical inference is important have been indicated by Stats; a T appears when the issue applies to theorists as well as to experimentalists. Few are the regions in which we can ignore statistics and statistical inference. Experiment design needs to consider from the start what statistic or summarized data form is required to achieve the desired outcome. There are then checks throughout the experiment, and finally there is analysis in which the measured statistics are used in inference.
1.2 Probability; probability distributions The concept of probability is crucial in decision processes, and there is a commonly accepted relationship between probability and statistics.
1.2 Probability; probability distributions
5
Table 1.1. Stages in astronomy experimentation Stage Observe
How
Examples
Considerations
Carefully
Experiment design: calibration, integration time Stats
What is wanted? Number of objects Stats
Reduce
Algorithms
Flat field Flux calibration
Data integrity Signal-to-noise T Stats
Analyse
Parameter estimation, Hypothesis testing T Stats
Intensity measurements Positions
Frequentist, Bayesian?
Hypothesistesting
Correlation tests Distribution tests
Conclude
T Reflect
Stats
Carefully
T
T
Stats
Stats
Mission achieved? A better way? ‘We need more data’ ? T Stats
T
Stats
Believable, Repeatable, Understandable? T Stats The next observations T
Stats
In a world in which our statistics are derived from finite amounts of data, we need probabilities as a basis for inference. For example, limited data yields only a partial idea of the point-spread function, such as the FWHM; we can only assign probabilities to the range of point-spread functions roughly matching this parameter. We all have an inbuilt sense of probability. We know for example that the height of adults is anything from say 1.5 to 2.5 metres. We know this from the totality of the population, all adults. But we know what a tall person is – and it is not necessarily somebody who is 2.5 m tall. The distribution is not flat; it peaks at around 1.7 m. The distribution of the heights of all adults, normalized to have an area of 1.0, is the measured probability density function, often called the probability distribution. (We meet them in a more rigorous context in Chapter 2.) The tails contain little area; and it is the tails that give us the decision: we probably call somebody tall when they are taller than 75 per cent of us.
6
Decision
We have made a decision based on a statistic, by relating that statistic to a probability distribution; we have decided that the person in question was tall. Note also what we did – observe, reduce, analyse, conclude, probably all in one glance. We did not do this rigorously in making a quantitative assessment of just how tall, which would have required a detailed knowledge of the distribution of height and a quantitative measurement. And reflect? Context of our observation? Why did we wish to register or decide that the person was tall? What next as a result? How was this person selected from the population? The brain has not only done the five steps but has also set the result into an extensive context; and this in processing the single glance. This is an example of a probability distribution for which there is unlikely to be a mathematical description, one determined by counting most of the population, or at least so much of it as to leave no doubt that it is well defined. There are distributions for which mathematical description is very precise, such as the Poisson and Gaussian (Normal) distributions, and there are many cases in which we have good reason to believe that these must represent the underlying distributions well. This is also an example of a ‘ruling-out’; here we ruled out the hypothesis that the person is of ‘ordinary’ height. There is a different type of statistical inference, the ‘ruling-in’ process, in which we compute the probability of getting a given result, and if it is ‘probable’, we accept the original hypothesis. It is also an example of ‘counting’ to find the probabilities, the frequency distribution. There are other ways of assigning probabilities, including opinion and states of knowledge; and in fact there are instances in which we are moderately comfortable with the paradoxical notion of assigning probabilities to unique events. It is essential that our view of statistics and statistical inference be broad enough to take such probability concepts on board. 1.3 Probability and statistics in inference What is the relationship between these two notions? Statistics, to anticipate later definitions, are combinations of the data that do not depend on any unknown parameters. The average is a common example. When we calculate the average of a set of data, we expect that it will bear some relation to the true, underlying mean of the distribution from which our data were drawn. In the classical tradition, we calculate the sampling distribution of the average, the probabilities of the various values it may assume as we (hypothetically) repeat our experiment many times. We
1.4 Non-parametric or distribution-free statistical inference
7
then know the probability that some range around our single measurement will contain the true mean. This is information that we can use to take decisions. This is precisely the utility of statistics – they are laboriously discovered combinations of observations which converge, for large sample sizes, to some underlying parameter we want to know (say, the mean). Useful statistics are actually rather few in number. There is another, radically different way of making inferences – the Bayesian approach. This focuses on the probabilities right away, without the intermediate step of statistics. In the Bayesian tradition, we invert the reasoning just described. The data, we say, are unique and known; it is the mean that is unknown, that should have probability attached to it. Without using statistics, we instead calculate the probability of various values of the mean, given the data we have. This also allows us to make decisions. In fact, as we shall see, this approach comes a great deal closer to answering the questions that scientists actually ask.
1.4 Non-parametric or distribution-free statistical inference There are four reasons why statistical inference based on known probability distributions does not work, or limits our possibilities severely. (i) We are measuring in experiments being run out there in the Universe, not by us. The underlying distributions may be far from known or understood; no averaging may be going on to lead us towards the central limit theorem and Gaussian distributions (see Chapter 2); yet we still wish to draw inferences about the underlying population. We only do so safely with non-parametric statistics, methods that do not require knowledge of the underlying distributions. (ii) We may have to deal with small-number samples, such as N = 3. Non-parametric techniques have the power to do this. (iii) The range of observation scales available to us is given in Table 1.2. Each such scale has a formal definition and formal properties. Each has admissible operations. Suffice it to say here that use of scales other than numerical (‘interval’) requires in most (but not all) cases that we use non-parametric methods. We may wish to make statistical inference without recourse to numerical scales. (iv) Others use such methods to draw inference. We need to understand what they are doing.
8
Decision Table 1.2. Measurement scales Scale type
Also called
Example and measurement
Nominal/ Categorical
Bins
Psychiatric types: schizophrenic, paranoid, manic-depressive, neurotic, psychopathic
Ordinal/ Ranking
Order
Army ranks: private, corporal, sergeant, major
Interval
Measures
Temperature: degrees Celsius
Non-parametric methods thus enormously increase the possibilities in decision-making and form an essential part of our process. They are described in the course of this book. 1.5 How to use this book This is not a textbook of statistical theory, a guide to numerical analysis, or a review of published work. It is a practical manual, which assumes that proofs, numerical methods and citation lists can easily be found elsewhere. This book sets out to tell it from an astronomer’s perspective, and our main objective is to help in gaining familiarity with the broad concepts of statistics and probability, to understand their usefulness, and to feel confident in applying them. Work through the examples and exercises; they are drawn from our experience and have been chosen to clarify the text. They vary in difficulty, from one-page calculations to miniprojects. Some need data; this may be simulated. If preferred, example datasets are available on the book’s website – as are the solutions to the exercises. Aim to become confident in the use of Monte Carlo simulations to check any calculations, and to try out ideas. Remember, in this subject we can do useful and revealing experiments – in the computer. Don’t be ashamed to let simulations guide your mathematical intuition! For further details on statistical methods and justification of theory, there is no substitute for a proper textbook. None of our topics is arcane and they will be found in the index of any elementary statistics book. We have found several particularly helpful: Mood, Graybill & Boes (1974), Lyons (1986), Barlow (1989), Lee (1997) and Bevington & Robinson (2002). Feigelson & Babu (1992a) and Babu & Feigelson (1996) cover many useful astronomical applications from a more rigorous point of view than we do.
Exercises
9
There is little algebra in this book; it would have greatly lengthened and cluttered the presentation to have worked through details. Likewise, we have not explained how various integrals were done or eigenvalues found. These things can be done by computers; packages such as the superb MATHEMATICA, used for many of the calculations in this book, can deal swiftly with more mathematical technology than most of us know. Using these packages frees us all up to think about the problem to hand, rather than searching in vain for missing minus signs or delving into handbooks for integrals which never seem to be there in quite the needed form. The other source is the indispensable Numerical Recipes (Press et al. 1992), which points the way for numerical solution of an enormous variety of problems, plus providing humorous and wise advice. We have not attempted exhaustive referencing. Rather, we have given enough key references to provide entry points to the literature. Online bibliographic databases provide excellent cross-referencing, showing who has cited a paper and who it cites; it is the work of minutes to collect a comprehensive reading list on any topic. The lecture notes for many excellent university courses are now on the Web; a well-phrased search may yield useful material to help with whatever is puzzling you. Finally, use this book as you need it. It can be read from front to back, or dipped into. Of course, no interesting topic is self-contained, but we hope the cross-referencing will connect all the technology needed to explore a particular topic. Exercises 1.1
1.2
At first sight, discovery of a new phenomenon may not read as an experiment as described in section 1.1. But it is. Describe the discovery of pulsars (Hewish et al. 1968) in terms of the five experimental stages. The significance of a certain conclusion depends very strongly on whether the most luminous known quasar is included in the dataset. The object is legitimately in the dataset in terms of prestated selection criteria. Is the conclusion robust? Believable?
2 Probability
God does not play dice with the Universe. (Albert Einstein)
Whether He does or not, the concepts of probability are important in astronomy for two reasons. (1) Astronomical measurements are subject to random measurement error, perhaps more so than most physical sciences because of our inability to rerun experiments and our perpetual wish to observe at the extreme limit of instrumental capability. We have to express these errors as precisely and usefully as we can. Thus when we say ‘an interval of 10−6 units, centred on the measured mass of the Moon, has a 95 per cent chance of containing the true value’, it is a much more quantitative statement than ‘the mass of the Moon is 1±10−6 units’. The second statement really only means anything because of some unspoken assumption about the distribution of errors. Knowing the error distribution allows us to assign a probability, or measure of confidence, to the answer. (2) The inability to do experiments on our subject matter leads us to draw conclusions by contrasting properties of controlled samples. These samples are often small and subject to uncertainty in the same way that a Gallup poll is subject to ‘sampling error’. In astronomy we draw conclusions such as ‘the distributions of luminosity in X-ray-selected Type I and Type II objects differ at the 95 per cent level of significance’. Very often the strength of this conclusion is dominated by the number of objects in the sample and is virtually unaffected by observational error. This chapter begins with a discussion of what probability is, and proceeds to introduce the concepts of conditionality and independence, providing a basis for the consequent discussion of Bayes’ theorem, with 10
2.1 What is probability?
11
prior and posterior probabilities. Only at this point is it safe to consider the concept of probability distributions; some common probability distributions are compared and contrasted. This sets the stage for the following chapter, dealing with statistics themselves, the penultimate product of data reduction – if conclusions/discoveries are considered as the ultimate product. The issues of expectation and errors, dependent on the distributions and statistics, are discussed in the final section of the following chapter.
2.1 What is probability? For a fascinating historical study of probability, see the books by Hald (1990; 1998). The ideas in this chapter draw heavily on the writings of Jaynes (1976; 1983; 2003). Another fundamental reference, rather heavy going, is Jeffreys (1961). The study of probability began with the analysis of games of chance involving cards or dice. Because of this background we often think of probabilities as a kind of limiting case of a frequency. Many textbook problems are still about dice, hands of cards, or coloured balls drawn from urns; in these cases it seems obvious to take the probabilities of certain events according to the ratio number of favourable events total number of events and the probability of throwing a six with one roll of the dice is ‘obviously’ 1/6. This probability derives from what Laplace called the ‘principle of indifference’, which in effect tells us to assign equal probabilities to events unless we have any information distinguishing them. In effect we have done the following calculation: probability of one spot = x probability of two spots = x probability of three spots = x and so on; this is the principle of indifference step. Further, we believe that we have identified all the cases; with the convention that the probability of a certain event (anything between one and six spots) is unity, we have 6x = 1.
12
Probability
This calculation, apparently trivial as it is, shows a vitally important feature: we cannot usefully define probability by this kind of ratio. We have had to assume that each face of the die is equally probable to start with – thus the definition of probability becomes circular. If we can identify equally likely cases, then calculating probabilities amounts simply to enumerating cases – not always easy, but straightforward in principle. However, identifying equally likely cases requires more thought. Many interesting and useful calculations can be done using the principle of indifference, either directly or by exploiting its applicability to aspects of the problem. For example, we may know that a die is biased, the faces are not ‘equally likely’. However, given some details of, say, the mass distribution of the die, we may be able to calculate the probabilities of the faces using an assumption that the initial direction of the throw is isotropic – in which case the principle of indifference applies to throw-directions. Sometimes we estimate probabilities from data. The probability of our precious observing run being clouded out is estimated by number of cloudy nights last year 365 but two issues arise. One is the limited data – we suspect that 10 years’ worth of data would give a different, more accurate result. The second issue is simply the identification of the ‘equally likely’ cases. Not all nights are equally likely to be cloudy, some student of these matters tells us; it’s much more likely to be cloudy in winter. What is ‘winter’, then? A set of nights equally likely to be cloudy? We can only estimate the probabilities correctly once we have identified the equally likely cases, and this identification is the subjective, intuitive step that is built into our reasoning about data from apparently malevolent instrumentation in an increasingly uncertain world. It is common to define probabilities as empirical statements about frequencies, in the limit of large numbers of cases – our 10 years’ worth of data. But, as we have seen, this definition must be circular because selecting the data depends on knowing which cases are equally likely. Defining probabilities in this way is sometimes called ‘frequentist’. It is sometimes the only way; but the risks must be recognized. So what is probability? The notion we adopt for the present is that probability is a numerical formalization of our degree or intensity of
2.1 What is probability?
13
belief. In everyday speech we often refer to the probability of unique events, showers of rain or election results. In the desiccated example of throwing dice, x measures the strength of our belief that any face will turn up. Provided that the die is not loaded, this belief is 1/6, the same for each face. Ascribing an apparently subjective meaning to probability in this way needs careful justification. After all, one person’s degree of belief is another person’s certainty, depending on what is known. We can only reason as best we can with the information we have; if our probabilities turn out to be wrong, the deficiency is in what we know, not the definition of probability. We just need to be sure that two people with the same information will arrive at the same probabilities. It turns out that this constraint, properly expressed, is enough to develop a theory of probability which is mathematically identical to the one often interpreted in frequentist terms. A useful set of properties of probability can be deduced by formalizing the ‘measure of belief’ idea. The argument is originally due to Cox (1946) and goes as follows: if A, B and C are three events and we wish to have some measure of how strongly we think each is likely to happen, then for consistent reasoning we should at least apply the rule if A is more likely than B, and B is more likely than C, then A is more likely than C. Remarkably, this is sufficient to put constraints on the probability function which are identical to the Kolmogorov axioms of probability, proposed some years before Cox’s paper: • Any random event A has a probability prob(A) between zero and one. • The sure event has prob(A) = 1. • If A and B are exclusive events, then prob(A or B) = prob(A) + prob(B). The Kolmogorov axioms are a sufficient foundation for the entire development of mathematical probability theory, by which we mean the apparatus for manipulating probabilities once we have assigned them.
Before 1987, four naked-eye supernovae had been recorded in ten centuries. What, before 1987, was the probability of a bright supernova happening in the twentieth century?
EXAMPLE
14
Probability
There are three possible answers. (1) Probability is meaningless in this context. Supernovae are physically determined events and when they are going to happen can, in principle, be accurately calculated. They are not random events. From this God’s-eye viewpoint, probability is indeed meaningless; events are either certain or forbidden. ‘God does not play dice . . . ’ (2) From a frequentist point of view our best estimate of the probability is 4/10, although it is obviously not very well determined. This assumes supernovae were equally likely to be reported throughout 10 centuries, which may well not be true. Eventually some degree of belief about detection efficiency will have to be made explicit in this kind of assignment. (3) We could try an a-priori assignment. In principle we might know the stellar mass function, the fate and lifetime as a function of mass, and the stellar birth rate. We would also need a detection efficiency. From this we could calculate the mean number of supernovae expected in 1987, and we would put some error bars around this number to reflect the fact that there will be variation caused by factors we do not know about – metallicity, perhaps, or location behind a dust cloud, and so on. The belief-measure structure is more complicated in this detailed model but it is still there. The model deals in populations, not individual stars, and assumes that certain groups of stars can be identified which are equally likely to explode at a certain time. Suppose now that we sight supernova 1987A. Is the probability of there being a supernova later in the twentieth century affected by this event? Approach (1) would say no – one supernova does not affect another. Approach (2), in which the probability simply reflects what we know, would revise the probability upward to 5/10. Approach (3) might need to adjust some aspects of its models in the light of fresh data; predicted probabilities would change.
Probabilities reflect what we know – they are not things with an existence all of their own. Even if we could define ‘random events’ (approach 1), we should not regard the probabilities as being properties of supernovae.
2.2 Conditionality and independence
15
2.2 Conditionality and independence Two events A and B are said to be independent if the probability of one is unaffected by what we may know about the other. In this case, it follows (not trivially!) from the Kolmogorov axioms that prob(A and B) = prob(A)prob(B).
(2.1)
Sometimes independence does not hold, so that we would also like to know the conditional probability: the probability of A, given that we know B. The definition is prob(A | B) =
prob(A and B) . prob(B)
(2.2)
If A and B are independent, knowing that B has happened should not affect our beliefs about the probability of A. Hence prob(A | B) = prob(A) and the definition reduces to prob(A and B) = prob(A)prob(B) again. If there are several possibilities for event B (label them B1 , B2 , . . .) then we have that prob(A) =
prob(A | Bi )prob(Bi ).
(2.3)
i
A might be a cosmological parameter of interest, while the Bs are not of interest. They might be instrumental parameters, for example. Knowing the probabilities prob(Bi ) we can get rid of these ‘nuisance parameters’ by a summation (or integration). This is called marginalization.
Take the familiar case in astronomy where some ‘remarkable’ event is observed, for example two quasars of very different redshifts close together on the sky. The temptation is to calculate an apriori probability, based on surface densities, of two specified objects being so close. However, the probability of the two quasars being close together is conditional on having noticed this fact in the first place. Thus the probability of the full event is simply prob(A | A) = 1, consistent with how we should expect to measure our belief in something that we
EXAMPLE
16
Probability
already know. We can say nothing further, although we might be able to formulate a hypothesis to carry out an experiment. Consider now the very different case in which we wish to know the probability of finding two objects of different types, say a galaxy and a quasar, within a specified angular distance r of each other. To be specific, we plan to search some fixed solid angle Ω. The surface densities in question are ςG and ςQ . On finding a galaxy, we will search around it for a quasar. We need prob(G in field and Q within r) = prob(Q within r | G in field)prob(G in field). This assumes that the probabilities are independent, obviously what we would like to test. A suitable model for the probabilities is the Poisson distribution (Section 2.4.2.2), and in the interesting case where the probabilities are small we have prob(G in field) = ςG Ω and prob(Q within r) = πr2 ςQ . The answer we require is therefore prob(G in field and Q within r) = ςG ςQ Ωπr2 . This is symmetrical in the quasar and galaxy surface densities as we would expect; it should not matter whether we searched first for a galaxy or for a quasar. Note the strong dependence on the search area that is specified before the experiment; if there is obscurity about this then the probabilities are not well determined.
As an extension of this example, it is possible to calculate the probability of finding triples of objects aligned to some small tolerance (Edmunds & George 1985). If the objects are all the same, the probability of a linear triple depends on the cube of the surface density and search area.
2.3 . . . and Bayes’ theorem
17
2.3 . . . and Bayes’ theorem 1
Bayes’ theorem is a simple equality, derived by equating prob (A and B) with prob (B and A). This gives the ‘theorem’: prob(B | A) =
prob(A | B)prob(B) . prob(A)
(2.4)
In this, the denominator is a normalizing factor. The theorem is particularly useful when interpreted as a rule for induction; the data, the event A, are regarded as succeeding B, the state of belief preceding the experiment. Thus prob(B) is the prior probability which will be modified by experience. This experience is expressed by the likelihood prob(A | B). Finally prob(B | A) is the posterior probability, the state of belief after the data have been analysed. Bayes’ theorem by itself is a perfectly innocent identity, a mathematical truism. It acquires its force from its interpretation. To see what this force is, we return to the familiar and simple problem of drawing those coloured balls from urns. It is clear, even automatic, what to calculate; if there are M red balls and N white balls, the probability of drawing three red balls and two white ones is . . . As a series of brilliant scientists realized, and as a series of brilliant scientists did not, this is generally not the problem we face. As scientists, we more often have a datum (three red balls, two white ones) and we are trying to infer something about the contents of the urn. This is sometimes called the problem of ‘inverse probability’. How does Bayes’ theorem help? We interpret it to be saying prob(contents of urn | data) ∝ prob(data | contents of urn) and of course we can calculate the right-hand side, given some assumptions. The urn example illustrates the principles involved; these are far more interesting than coloured balls. 1
Who was Bayes? Thomas Bayes (1702–61) was an English vicar, mathematician and statistician. His bibliography consisted of three works: one (by the vicar) on divine providence, the second (by the mathematician) a defence of the logical bases of Newton’s calculus against the attacks of Bishop Berkeley, and the third (by the statistician and published posthumously) the famous Essay Towards Solving a Problem in the Doctrine of Chances. There is speculation that it was published posthumously because of the controversy which Bayes believed would ensue. This must be an a-posteriori judgement. Surely Bayes could never have imagined the extent of this controversy without envisaging the nature and extent of modern scientific data.
18
Probability
There are N red balls and M white balls in an urn; we know the total N + M = 10, say. We draw T = 3 times (putting the balls back after drawing them) and get R = 2 red balls. How many red balls are there in the urn? Our model (hypothesis) is that the probability of a red ball is N . N +M We assume that the balls are not stratified, arranged in pairs, or anything else ‘peculiar’. The probability of getting R red balls, the likelihood, is R T −R N M T . R N +M N +M
EXAMPLE
This is the number of permutations of the R red balls amongst the T draws, multiplied by the probability that R balls will be red and T − R will not be red. (This is a Binomial distribution; see section 2.4.2.1.) Thus we have the probability (data, given the model) part of the righthand side of Bayes’ theorem. We also need probability (model), or the prior. We assume that the only uncertain bit of the model is N , which to start with we take as being uniformly likely between zero and N + M . Without bothering with the details at the moment, we plot up the lefthand side of Bayes’ theorem (the posterior probability) as a function of N – see Fig. 2.1. For a draw of, say, three red balls in five tries, the posterior probability peaks at 6; for 30 out of 50, the peak is still at 6 but other possibilities are much less likely.
Probability
0.5 0.4 0.3 0.2 0.1 0 0
2
4 6 8 True number of red balls
10
Fig. 2.1. The probability distribution of the number of red balls in the urn, for five (solid curve) and 50 drawings (dashed curve).
2.3 . . . and Bayes’ theorem
19
This seems unsurprising and in accord with common sense – but notice that we are speaking now of the probability of there being 1, 2, 3, . . . red balls in a unique urn that is the subject of our experiment. We are describing our state of belief about the contents of the urn, given what we know (the data, and our prior information). The key point of this example is that we have succeeded in answering our scientific question: we have made an inference about the contents of the urn, and can make probabilistic statements about this inference. For example, the probability of the urn containing three or fewer red balls is 11 per cent. We are assigning probabilities to these statements to N because we are using probability to reflect our degree of certainty. Our concern, as experimental scientists, is with what we can infer about the world from what we know. Bayes’ theorem allows us to make inferences from data, rather than compute the data we would get if we happened to know all the relevant information about our problem. This may seem academic; but suppose we had data from two populations and wanted to know if the means were different. Many chapters of statistics textbooks answer the opposite question for us: given populations with two different means, what data would you get? The combination of interpreting probability as a consistent measure of belief, plus Bayes’ theorem, allows us to answer the question we wish to pose: given the data, what are the probabilities of the parameters contained in our statistical model? Another very significant point about this example is the use of prior information; again, we assigned probabilities to N to reflect what we know. Notice that although the word ‘prior’ suggests ‘before the experiment’ it really means ‘what we know apart from the data’. Sometimes this can have a dramatic, even disconcerting effect on our inferences:
Suppose we make an observation with a radio telescope at a randomly selected position in the sky. Our model of the data (an event labelled D, consisting of the single measured flux density f ) is that it is distributed in a Gaussian way (Section 2.4.2.3) about the true flux density S with a variance (Section 2.4.2) σ 2 . The extensive body of radio source counts also tells us the a-priori distribution of S; for the purposes of this example, we approximate this information by the simple prior
EXAMPLE
prob(S) = KS −5/2
20
Probability
describing our prior state of knowledge. K normalizes the counts to unity; there is presumed to be one source in the beam at some fluxdensity level. The probability of observing f when the true value is S we take to be 1 exp − 2 (f − S)2 . 2σ Bayes’ theorem then tells us 1 2 prob(S | D) = K exp − 2 (f − S) S −5/2 , 2σ
with the normalizations condensed into the single parameter K . If we were able to obtain n independent flux measurements fi then the result would be n 1 2 prob(S | D) = K exp − 2 (fi − S) S −5/2 . 2σ i=1 Suppose, for specific example, that the source counts were known to extend from 1 to 100 units, the noise level was σ = 1, and the data were 2, 1.3, 3, 1.5, 2 and 1.8. In Fig. 2.2 are the posterior probabilities for the first two, then four, then six measurements. The increase in data gradually overwhelms the prior but the prior affects conclusions markedly (as it should) when there are few measurements.
Probability density
1.4 1.2 1 0.8 0.6 0.4 0.2 0 1
1.5
2
2.5 3 Flux density
3.5
4
Fig. 2.2. Measurement of flux density given a power-law prior (source count) and a Gaussian error distribution. The posterior probability distribution for flux density is plotted for two, four and then six of the measurements listed in the text; the form of the curve approaches Gaussian as numbers increase.
2.3 . . . and Bayes’ theorem
21
If subsequently we looked at a survey plate of the region we had observed, and found that the radio emission was from some category of object (say, a quasar) with different source counts, our prior would change and so would the posterior probability. In turn, our idea of the most probable flux density would also change.
In this example, the prior seems to be well determined. However, in some cases we wish to estimate quantities where the argument is not so straightforward. What would we take as the prior in the previous example if we were making the first ever radio measurements? Or if we needed an estimate of the mean of a Gaussian, then we have to ask how we interpret the prior probability of the mean. Sometimes we even need a probability of a probability:
Return to the question of supernova rate per century and consider how to estimate this; call this ρ. Our data are four supernovae in ten centuries. Our prior on ρ, expressing our total ignorance, is uniform between 0 and 1; we have no preconceptions or information about ρ. A suitable model for prob(data | ρ) is the Binomial distribution (Section 2.4.2.1), because in any century we either get a supernova or we do not (neglecting here the possibility of two supernovae in a century). Our posterior probability is then 10 prob(ρ | data) ∝ ρ4 (1 − ρ)6 × prior on ρ. 4
EXAMPLE
We follow Bayes and Laplace in taking the prior to be uniform in the range 0 to 1. Then, to normalize the posterior probability properly we need 1 prob(ρ | data) dρ = 1, 0
resulting in the normalizing constant 1 10 ρ4 (1 − ρ)6 dρ, 4 0 which happens to be Γ(10)Γ(4) = B[5, 7], Γ(14)
22
Probability
Probability density
2.5 2 1.5 1 0.5 0 0
0.2 0.4 0.6 0.8 Number of supernovae per century
1
Fig. 2.3. The posterior probability distribution for ρ, given that we have four supernovae in ten centuries.
where B is the (tabulated) beta function. In general, for n supernovae in m centuries, the distribution is prob(ρ | data) =
ρn (1 − ρ)m−n . B[n + 1, m − n + 1]
Our distribution (n = 4, m = 10) peaks – unsurprisingly – at 4/10, as shown in Fig. 2.3.
As the sample size increases the distribution becomes narrower so that the peak posterior probability is more and more closely defined by the ratio of successes (supernovae, in our example) to sample size. This result is sometimes called the law of large numbers, expressing as it does the frequentist idea of a large number of repetitions resulting in a converging estimate of probability. The key step in this example is ascribing a probability distribution to ρ, in itself a probability. This makes no sense in a frequentist approach, nor indeed in any interpretation of probabilities as objective. Even if we are prepared to leap this metaphysical hurdle, in very many cases the assignment of a prior probability is much more difficult than in this example. Indeed, it is certain that the assignment of priors in the current example has been greatly oversimplified. Both Jeffreys (1961) and Jaynes (1968) discuss the prior on ρ, arguing that in many cases a uniform prior is far too agnostic. By intricate
2.3 . . . and Bayes’ theorem
23
arguments, they arrive at other possibilities: prob(ρ) =
1 ρ(1 − ρ)
and the ‘Haldane prior’ 1
prob(ρ) =
ρ(1 − ρ)
.
These are intended to reflect the fact that in most experiments we are expecting a yes or no answer. Assigning priors when our knowledge is rather vague can be quite difficult, and there has been a long debate about this. Some ‘obvious’ priors (such as the one we might use for location, simply uniform from −∞ to ∞) are not normalizable and can sometimes get us into trouble. Out of the enormous literature on this subject, try Lee (1997) for an introduction, and Jaynes’s writings for some fascinating arguments. One of the ways of determining a prior is the maximum entropy principle; we will see an example of such a prior later (Section 6.7). A common prior for a scale factor σ is Jeffrey’s prior, uniform in log σ.
Finally, the use of Bayes’ theorem as a method of induction can be neatly illustrated by our supernova example. For simplicity, imagine that we establish our posterior distribution at the end of the nineteenth century, so that it is ρ4 (1 − ρ)6 /B[5, 7], as shown earlier. At this stage, our data are four supernovae in ten centuries. Reviewing the situation at the end of the twentieth century, we take this as our prior. The available new data consist of one supernova, so that the likelihood is simply the probability of observing exactly one event of probability ρ, namely ρ. The updated posterior distribution is
EXAMPLE
prob(ρ | data) =
ρ5 (1 − ρ)6 B[6, 7]
which peaks at ρ = 5/11 as we might expect.
In these examples we have focused on the peak of the posterior probability distribution. This is one way amongst many of attempting to characterize the distribution by a single number. Another choice is the
24
Probability
posterior mean, defined by
1
ρ prob(ρ | data) dρ.
=
(2.5)
0
If we have had N successes and M failures, the posterior mean is given by a famous result called Laplace’s rule of succession: =
N +1 . N +M +2
In our example, at the end of the nineteenth century Laplace’s rule would give 5/12 as an estimate of the probability of a supernova during the twentieth century. This differs from the 4/10 derived from the peak of the posterior probability, and it will do so in general. Unless posterior distributions are very narrow, attempting to characterize them by a single number is frequently misleading. How best to characterize the distribution depends on what is to be done with the answer, which in turn depends on having a carefully posed question in the first place.
2.4 Probability distributions 2.4.1 Concept We have referred several times to probability distributions. The basic idea is intuitive; here is a little more detail. Consider the fascinating experiment in which we toss four ‘fair’ coins. The probability of no heads is (1/2)4 ; of one head 4×(1/2)4 ; of two heads 6 × (1/2)4 , etc. The sum of the possibilities for getting no heads to four heads is readily seen to be 1.0. If x is the number of heads (0, 1, 2, 3, 4), we have a set of probabilities prob(x) = (1/16, 1/4, 3/8, 1/4, 1/16); we have a probability distribution, describing the expectation of occurrence of event x. This probability distribution is discrete; there is a discrete set of outcomes and so a discrete set of probabilities for those outcomes. In this sort of case we have a mapping between the outcomes of the experiment and a set of integers. Sometimes the set of outcomes maps onto real numbers instead, the set of outcomes no longer containing discrete elements. We deal with this by the contrivance of discretizing the range of real numbers into little ranges within which we assume the probability does not change. Thus if x is the real number that indexes outcomes, we associate with it a probability density f (x); the
2.4 Probability distributions
25
probability that we will get a number ‘near’ x, say within a tiny range δx, is prob(x) δx. We loosely refer to probability ‘distributions’ whether we are dealing with discrete outcomes or not. Formally: if x is a continuous random variable, then f (x) is its probability density function, commonly termed probability distribution, when
b (i) prob(a < x < b) = a f (x) dx, ∞ (ii) −∞ f (x) dx = 1, and (iii) f (x) is a single-valued non-negative number for all real x.
xThe corresponding cumulative distribution function is F (x) = f (y) dy. Probability distributions and distribution functions may −∞ be similarly defined for sets of discrete values of x; and distributions may be multivariate, functions of more than one variable.
2.4.2 Some common distributions The better-known probability density functions appear in Table 2.1 together with location (where is the ‘centre’ ?) and dispersion (what is the ‘spread’ ?) quantifiers. These quantifiers can be given by the first two moments of the distributions (Section 3.1): ∞ µ1 (mean) = µ = xf (x) dx (2.6) −∞ ∞ µ2 (variance) = σ 2 = (x − µ1 )2 f (x) dx. (2.7) −∞
σ is known as the standard deviation. Three of them are of prime importance, the Binomial, Poisson, and Gaussian or Normal, and we discuss these in turn. 2.4.2.1 Binomial distribution There are two outcomes – ‘success’ or ‘failure’. This common distribution gives the chance of n successes in N trials, where the probability of a success at each trial is the same, namely ρ, and successive trials are independent. This probability is then N prob(n) = ρn (1 − ρ)N −n . (2.8) n
Student t
Chi-square
Normal (Gaussian)
Poisson
Binomial
Uniform
Distribution
1 √ 2π
exp −
0
µ
(1+t2 /ν)−[(ν+1)/2] √ πνΓ(ν/2)
ν
(x−µ)2 2σ 2
µ
np
(a + b)/2
Mean
exp(− 12 χ2 )
χ2(ν/2−1) 2ν/2 Γ(v/2)
σ
f (t; ν) = Γ[(ν + 1)/2]
f (χ2 ; ν) =
f (x; µ, σ) =
e−µ µx x!
n! px q n−x x!(n−x)!
f (x; µ) =
f (x; p, q) =
f (x; a, b) = 1/(b − a) a < x < b = 0, x < a, x > b
Density function
Table 2.1. The common probability density functions
Vital in the comparison of samples, model testing; characterizes the dispersion of observed samples from the expected dispersion, because if xi is a sample of ν variables Normally and independently distributed with means µi and 2 2 2 variances σi2 , then χ2 = ΣN i=1 (xi − µi ) /σi obeys f (χ ; ν). Invariably tabulated and used in integral form. Tends to the Normal distribution as ν → ∞. For comparison of means, Normally distributed populations; (µ, σ), and if if n xi ’s are taken from a Normal population √ xs and σs are determined, then t = n(xs − µ)/σs is distributed as f (t, ν) where the ‘degrees of freedom’ ν = n − 1. The statistic t can also be formulated to compare means for samples from Normal populations with the same σ, different µ. Tends to Normal as ν → ∞.
ν/(ν − 2) (for ν > 2)
The essential distribution; see text. The central limit theorem ensures that the majority of ‘scattered things’ are dispersed according to f (x; µ, σ).
The limit for the Binomial distribution as p 1, setting µ ≡ np. It is the ‘count-rate’ distribution, e.g. take a star from which an average of µ photons are received per ∆t (out of a total of n emitted; hence p 1); the probability of receiving x photons in ∆t is f (x; µ). Tends to the Normal distribution as µ → ∞.
x is the number of ‘successes’ in an experiment with two possible outcomes, one (‘success’) of probability p, and the other (‘failure’) of probability q = 1 − p. Becomes a Normal distribution as n → ∞.
In the study of rounding errors; as a tool in studies of other continuous distributions.
Raison d’ˆ etre
2ν
σ2
µ
npq
(b − a)/12
Variance
2.4 Probability distributions
27
The leading term, the combinatorial coefficient, gives the number of distinct ways of choosing n items out of N : N! N = . (2.9) n n!(N − n)! This coefficient can be derived in the following way. There are N ! equivalent ways of arranging the N trials. However there are n! permutations of the successes, and (N − n)! permutations of the failures, which correspond to the same result – namely, exactly n successes, arrangement unspecified. Since we require not just n successes (probability pn ) but exactly n successes, we need exactly N −n failures, probability (1−p)(N −n) as well. The Binomial distribution follows from this argument. The Binomial distribution has a mean value given by N
n prob(n) = N p
n=0
and a variance or mean square value of N
(n − N p)2 prob(n) = N p(1 − p).
n=0
Suppose we know, from a sample of 100 galaxy clusters selected by automatic pattern-recognition techniques, that ten contain a dominant central galaxy. We plan to check a different sample of 30 clusters, now selected by X-ray emission. How many of these clusters do we expect to have a dominant central galaxy? If we assume that the 10 per cent probability holds for the X-ray sample, then the chance of getting n dominant central galaxies is 30 prob(n) = 0.1n 0.930−n . n
EXAMPLE
For example, the chance of getting 10 is about 1 per cent; if we found this many we would be suspicious that the X-ray cluster population differed from the general population. Suppose we made these observations and did find 10 centrally dominated clusters. What can we do with this information? The Bayesian thing to do is a calculation that parallels the supernova example. Assuming the X-ray galaxies are a homogeneous set, we can
28
Probability
Probability density
12 10 8 6 4 2 0 0
0.2 0.4 0.6 0.8 Fraction centrally dominated
1
Fig. 2.4. The posterior probability distribution for the fraction of X-rayselected clusters that are centrally dominated. The black line uses a uniform prior distribution for the fraction; the dashed line uses the prior derived from an assumed previous sample in which 10 out of 100 clusters had dominant central members. The light curve shows the distribution for this earlier sample.
deduce the probability distribution for the fraction of these galaxies that have a dominant central galaxy. A relevant prior would be the results for the original larger survey. Figure 2.4 shows the results, making clear that the data are not really sufficient to alter our prior very much. For example, there is only a 10 per cent chance that the centrally dominant fraction exceeds even 0.2; and indeed Fig. 2.4 shows that the possibility of it being as high as 33 per cent is completely negligible. Our X-ray clusters have a different prior from the general population.
The Binomial distribution is the parent of two other famous distributions, the Poisson and the Gaussian. 2.4.2.2 Poisson distribution The Poisson distribution derives from the Binomial in the limiting case of very rare events and a large number of trials, so that although p → 0, N p → a finite value. Calling the finite mean value µ1 = µ, the Poisson distribution is µn −µ prob(n) = (2.10) e . n! The variance of the Poisson distribution, µ2 , is also µ.
2.4 Probability distributions
29
A familiar example of a process obeying Poisson statistics is the number of photons arriving during an integration. The probability of a photon arriving in a fixed interval of time is (often) small. The arrivals of successive photons are independent (apart from small correlations arising because photons obey Bose–Einstein statistics, negligible for our purposes). Thus the conditions necessary for the Poisson distribution are met. Hence, if the integration over time t of photons arriving at a rate λ has a mean of µ = λt photons, then the fluctuation on this √ number will be σ = µ. (In practice we usually only know the number of photons in a single exposure, rather than the mean number; obviously we can then only estimate the µ. This case is the subject of an exercise in the next chapter.) For photon-limited √ observations, such as CCD images or spectra, µ = λt while σ = λt. If we ‘integrate’ more, √ σ ∝ t, while signal ∝ t. √ Thus Signal/Noise ∝ t, the sky-limited case. There are the following further cases:
EXAMPLE
(i) Photon-limited, e.g. CCD observations of faint objects: √ µ S/N ∝ √ , or ∝ t. µ (ii) Readout-limited, e.g. CCD observations of bright objects: S/N ∝
µ σccd
, or
∝t
for CCD of readout noise σccd . (iii) Receiver-limited, e.g. radio astronomy: S/N ∝
S √ , or σrec / t
∝
√
t
for a receiver of thermal noise σrec .
2.4.2.3 Gaussian (Normal) distribution Both the Binomial and the Poisson distributions tend to the Gaussian distribution (Fig. 2.5), large N in the case of the Binomial, large µ in
30
Probability
Probability density
0.4 0.3 0.2 0.1 0
−3
−2
−1
0 x
1
2
3
Fig. 2.5. The Normal (Gaussian) distribution. The area under the curve is 1.00; the area between ±1σ is 0.68; between ±2σ is 0.95; and between ±3σ is 0.997.
the case of the Poisson. The (univariate) Gaussian (Normal) distribution is 1 1 prob(x) = √ exp − 2 (x − µ)2 (2.11) σ 2π 2σ from which it is easy to show that the mean is µ and the variance is σ 2 (Section 3.1). How this comes about for the Binomial distribution is the subject of an exercise. For the Binomial when the sample size is very large, the discrete distribution tends to a continuous probability density 1 1 2 prob(n) = √ exp − 2 (n − µ) σ 2π 2σ in which the mean µ = N p and variance σ 2 = N p(1 − p) are still given by the parent formulae for the Binomial distribution. Here is an instance of the discrete changing to the continuous distribution: in this approximation we can treat n as a continuous variable (because n changes by one unit at a time, being an integer, and so the fractional change 1/n is small). The true importance of the Gaussian distribution and its dominant position in experimental science, however, stems from the central limit theorem. A non-rigorous statement of this is as follows.
2.4 Probability distributions
31
Form averages Mn from repeatedly drawing n samples from a population with finite mean µ, variance σ 2 . Then the distribution of (Mn − µ) √ → Gaussian distribution σ/ n with mean 0, variance 1, as n → ∞. This is a remarkable theorem. What it says is that provided certain conditions are met – and they are in almost all physical situations – a little bit of averaging will produce a Gaussian distribution of results no matter what the shape of the distribution from which the sample is drawn. Even eyeball integration counts. It means that errors on averaged samples will always look ‘Gaussian’. The reliance on Gaussian distributions, made valid by the unsung hero of statistical theory and indeed experimentation, the central limit theorem, shapes our entire view of experimentation. It is this theorem which leads us to describe our errors in the universal language of sigmas, and indeed to argue our results in terms of sigmas as well, which we explicitly or implicitly recognize as describing our place within or at the extremities of the Gaussian distribution. Figure 2.6 demonstrates the compelling power of the central limit theorem. Here we have brutally truncated an exponential, clearly an extremely non-Gaussian distribution. The histogram obtained in drawing 200 random samples from the distribution follows it closely. When 200 values resulting from averaging just four values have been formed, the distribution is already becoming symmetrical; by the time 200 values of 16 long averages have been formed, it is virtually Gaussian. Before leaving the central limit miracle and Gaussian distributions, it is important to emphasize how tight the tails of the Gaussian distribution are (Table A2.2). The range ±2σ encompasses 95.45 per cent of the area. Thus the infamous 2σ result has a less than 5 per cent chance of occurring by chance. But we scoff – because the error estimates are difficult to make, and observers are optimistic. Things upset the distribution; there are outlying points. Thus astronomers feel it necessary to quote results in the range 3σ to even 10σ, casting inevitable doubt on belief in their own error estimates. In fact, experimentalists are aware of another key feature of the central limit theorem: the convergence to a Gaussian happens fastest at the centre of the distribution, but the wings may converge much more slowly to a Gaussian form. Interesting results (the 10σ ones) of course acquire their probabilistic interpretation from knowing the shape of the tails to high accuracy.
Probability
y
y
x
x
32
x
y
y
x
x
x
x
x
Fig. 2.6. An indication of the power of the central limit theorem. The panels show successive amounts of ‘integration’: in the upper left panel, a single value has been drawn; in the upper right, 200 values have been formed from an average of two values; lower left, 200 values from an average of four; lower right, 200 values from an average of 16.
2.5 Inferences with probability What can we do with Bayesian probability calculations? We will use these many times in the rest of this book, but here is a summary of the method. First, we may estimate parameters. This is closely related to the field of data modelling (Section 6.1). We have a probability distribution f (data | α ) and we wish to know the parameter vector α . The Bayesian route is clear; compute the posterior distribution of α , as we have shown in several examples in this chapter.
Suppose we have N data Xi , drawn from a Gaussian of known variance σ 2 but unknown mean µ. The parameter we want is µ. To proceed, we need a prior on µ; we take the so-called ‘diffuse’ prior,
EXAMPLE
2.5 Inferences with probability
33
where prob(µ) = constant over some wide range of µ, the range defined by our knowledge of the problem. Of course we might have more precise information available. From Bayes, the posterior distribution follows at once: N 2 i=1 (Xi − µ) f (µ | data) ∝ exp − 2σ 2 and with some simplification we get f (µ | data) ∝ exp −
1 N
N
2 i=1 (Xi − µ)
2
2 σN
so that the average of the data is distributed around µ, with variance σ 2 /N . One of the exercises is to find the distribution of the variance, knowing the mean.
This method is related to the classical technique of maximum likelihood. If the prior is ‘diffuse’, as in the example, then the posterior probability is proportional to the likelihood term f (data | α ). Maximum likelihood picks out the mode of the posterior, the value of α which maximizes the likelihood. This amounts to characterizing the posterior by one number, an approach which is often useful because of powerful theorems on maximum likelihood. We consider this in more detail in Section 6.1; some exercises at the end of this chapter illustrate the procedure. Often knowing the posterior distribution of the parameter of interest is enough; we might be making a comparison with an exactly known quantity, perhaps derived from some theory. However we may wish to compare with an experimental determination of some other parameter A typical case, for scalar parameters α and β, would be to ask for the β. probability that, say, α is bigger than β. Suppose therefore we have derived two distributions prob(α) = pA (α) and prob(β) = pB (β) from independent samples. The probability that α is larger than β is ∞ ∞ p(α > β) = pB (y) dy pA (x) dx −∞
y
34
Probability
and the double integral simplifies to ∞ p(α > β) = (1 − CA (x)pB (x) dx −∞
in which CA is the cumulative distribution corresponding to pA . If pA and pB are the same distribution, this becomes p(α > β) = 1/2, as expected. Usually these integrals have to be done numerically case by case, but are worth the effort. We may express posterior probabilities by using the notion of odds, a handy way of expressing probabilities when we have only two possibilities. The odds on event A are just prob(A) . prob(not A) For instance, the odds on throwing a 6 with a fair die are 5 to 1 (probability of 1/6 for throwing a six, 5/6 for anything else). From a betting point of view, the odds on a bet give the profit that might be made on a stake; in the case of our example with dice, being offered 5 to 1 odds for a 6 means we would get $5 profit ($6 payout) on a stake of $1, if a 6 comes up. Of course a bookie will offer slightly different odds, to be sure of a profit in the long run. If we have two exclusive possibilities for a prior, say A and not A, then the posterior odds are given by the ratio of the posterior probabilities with each prior, and give an indication of which prior to bet on, given the available data.
Exercises 2.1
A warm-up on coin-tossing. This is not an astronomical problem but does provide a warm-up exercise on probability and random numbers. Every computer has a way of producing a random number between zero and one. Use this to simulate a simple coin-tossing game where player A gets a point for heads, player B a point for tails. Guess how often in a game of N tosses the lead will change; if A is in the lead at toss N , when was the previous change of lead most likely to be? And by how much is a player typically in the lead? Try to back these guesses up with calculations, and then simulate the game. For many more game-based illustrations of probability, see Haigh (1999).
Exercises 2.2
2.3
2.4
2.5
2.6
2.7
35
Efficient choosing. Imagine you are on a 10-night observing run with a colleague, in settled weather. You have an agreement that one of the nights, of your choosing, will be for your exclusive use. Show that, if you wait for five nights and then choose the first night that is better than any of the five, you have about a 25 per cent chance of getting the best night of the ten. For a somewhat harder challenge, find the optimum length of the ‘training sample’. Bayesian inference. Consider the proverbial bad penny, for which prior information has indicated that there is a probability of 0.99 that it is unbiased (‘ok’); or a probability of 0.01 that it is double-headed (‘dh’). What is the (Bayesian) posterior probability, given this information, of obtaining seven heads in a row? In such a circumstance, how might we consider the fairness of the coin? Or of the experimenter who provided us with the prior information? What are the odds on the penny being fair? Laplace’s rule and priors. Laplace’s rule (Section 2.3) ρ = (N + 1)/(N + M + 2) depends on our prior for ρ. If we have one success and no failures, consider what the rule implies, and discuss why this is odd. How is the rule changed for alternative priors, for example Haldane’s? Bayesian reasoning in an everyday situation. The probability of a certain medical test being positive is 90 per cent, if the patient has disease D. If your doctor tells you the test is positive, what are your chances of having the disease? If your doctor also tells you that 1 per cent of the population have the disease, and that the test will record a false positive 10 per cent of the time, use Bayes’ theorem to calculate the chance of having D if the test is positive. Inverse Chi-squared statistic. For a Gaussian of known mean (say zero), show that the posterior distribution for the variance is “inverse” χ2 . Use the ‘Jeffreys prior’ for the variance: prob(σ) = 1/σ. Comment on the differences between this result, and the one obtained by using a uniform prior on σ. Maximum likelihood and the Poisson distribution. Suppose we have data which obey a Poisson distribution with parameter µ, and in successive identical intervals we observe n1 , n2 , . . . events. Form the likelihood function by taking the product
36
2.8
2.9
Probability of the distributions for each ni , and differentiate to find the maximum-likelihood estimate of µ. Is it what you expect? Maximum likelihood and the exponential distribution. Suppose we have data X1 , X2 , . . . from the distribution 1/2a exp(−|x|/a). Compute the posterior distribution of a for a uniform prior, and Jeffreys’s prior prob(a) ∝ 1/a. Do the differences seem reasonable? Which prior would you choose? If a were known, but the location µ was to be found, what would be the maximum-likelihood estimate? Birth control. Imagine a society where boys and girls were (biologically) equally likely to be born, but families cease producing children after the birth of the first boy. Are there more males than females in the population? Attack the problem in three ways: pure thought, by a simulation, and by an analytic calculation.
3 Statistics and expectations
Lies, damned lies and statistics. (Benjamin Disraeli)
In embarking on statistics we are entering a vast area, enormously developed for the Gaussian distribution in particular. This is classical territory; historically, statistics were developed because the approach now called Bayesian had fallen out of favour. Hence direct probabilistic inferences were superseded by the indirect and conceptually different route, going through statistics and intimately linked to hypothesis testing. The use of statistics is not particularly easy. The alternatives to Bayesian methods are subtle and not very obvious; they are also associated with some fairly formidable mathematical machinery. We will avoid this, presenting only results and showing the use of statistics, while trying to make clear the conceptual foundations.
3.1 Statistics Statistics are designed to summarize, reduce or describe data. The formal definition of a statistic is that it is some function of the data alone. For a set of data X1 , X2 , . . . , some examples of statistics might be the average, the maximum value or the average of the cosines. Statistics are therefore combinations of finite amounts of data. In the following discussion, and indeed throughout, we try to distinguish particular fixed values of the data, and functions of the data alone, by upper case (except for Greek letters). Possible values, being variables, we will denote in the usual algebraic spirit by lower case.
37
38
Statistics and expectations
The summarizing aspect of statistics is exemplified by those describing (1) location and (2) spread or scatter. (1) The location of the data can be indicated by various combinations: N Average, denoted by overlining: X = 1/N i=1 Xi . Median: arrange Xi according to size; renumber. Then Xmed = Xj where j = N/2 + 0.5, N odd, Xmed = 0.5(Xj + Xj+1 ) where j = N/2, N even. Mode: Xmode is the value of xi occurring most frequently; it is the location of the peak in the histogram of Xi . (2) Statistics indicating the scale or amount of scatter in the data are, for example, N Mean deviation: ∆X = (1/N ) i=1 |Xi − X|. N Mean square deviation: S 2 = (1/N ) i=1 (Xi − X)2 . Root mean square deviation: rms = S. We are so familiar with statistics like these that a result such as ‘D = 8.3 ± 0.1 Mpc’ provokes no questions. But what does it mean? It does not tell us the probability that the true value of D is between 8.2 and 8.4. We usually assume that a Gaussian distribution applies, placing our faith in the central limit theorem. Knowing the distribution of the errors allows us to make probabilistic statements, which are what we need. After all, if there were only a 1 per cent chance that the interval [8.2, 8.4] contained the true value of D, we might not regard the stated error as being very useful. So this is one key aspect of statistics; they are associated with distributions. In fact they are most useful when they are estimators of the parameters of distributions. In quoting our measurement of D, we are hoping that 8.3 is an estimate of the parameter µ of some Gaussian, while 0.1 is an estimate of σ. The other key aspect of statistics is that they are to be interpreted in a classical, not Bayesian framework. We need to look carefully at this distinction; it parallels our discussion of those coloured balls in the urn. Assuming a true distance D0 , a classical analysis tells us that D is (say) Normally distributed around D0 , with a standard deviation of 0.1. So we are to imagine many repetitions of our experiment, each yielding a value of the estimate D which dances around D0 . We might form a confidence interval (such as [8.2, 8.4]) which will also dance around randomly, but will contain D0 with a probability we can calculate. Just as in the case of the coloured balls, this approach assumes the thing we want to know, and tells us how the data will behave.
3.1 Statistics
39
A Bayesian approach circumvents all this; it deduces directly the probability distribution of D0 from the data. It assumes the data, and tells us the thing we want to know. There are no imagined repetitions of the experiment. Conceptually it is clearer than classical methods, but these are so well developed and established (particularly for the Gaussian) that we will give some explanation of classical statistics now, and indeed use classical results in many places in this book. It is worth remembering, however, that statistics of known usefulness are quite rare; the intensive development of statistics based on the Gaussian should not blind us to this fact. In many cases of astronomical interest we may need to derive useful statistics for ourselves. By far the easiest method for doing this is maximum likelihood (Section 6.1) and this is so close to a Bayesian method that we may expect to be doing Bayesian, not classical, inference in any new case where we cannot draw immediately on classical results. To repeat, statistics are properties of the data and only of the data; they summarize, reduce, or describe the data. Variables such as µ and σ of the Poisson and Gaussian distributions define these distribution functions and are not statistics. But we may anticipate that our data do follow these or other distributions and we may therefore wish to relate statistics from the data to parameters describing the distributions. This is done through expectations or expectation values, long-run average properties depending on distribution functions. The expectation E[f (x)] of some function f of a random variable x, with distribution function g, is defined as E[f (x)] = f (x)g(x) dx (3.1) i.e. the sum of all possible values of f , weighted by the probability of their occurrence. We can think of the expectation as being the result of repeating an experiment many times, and averaging the results. We might, for example, compute an average X; if we repeat the experiment many times, we will find that the average of X will converge to the true mean value, the expectation of the function f (x) = x: E[x] = xg(x) dx. (3.2) Note that the expectation is not to be understood as referring to a very large sample; we can ask for the expectation value of a combination of a finite number of data.
40
Statistics and expectations The statistic S 2 should likewise converge to the variance, defined by var[x] = E[(x − µ)2 ] = (x − µ)2 g(x) dx.
(3.3) (3.4)
However, as we shall see, we do have to take some care that the integrals actually exist.
Take our favourite distribution, the Gaussian. The probability density of getting a datum x near µ is 1 −(x − µ)2 g(x | µ, σ) = √ exp 2σ 2 2πσ
EXAMPLE
but what are these parameters µ and σ? It’s not difficult to show (changing variables and using standard identities) that E[x] = xg(x | µ, σ) dx = µ, (3.5) and E[(x − µ)2 ] =
(x − µ)2 g(x | µ, σ) dx = σ 2 .
(3.6)
We would therefore expect that the average X and mean square deviation S 2 would be related to µ and σ 2 . As any statistics text will show, indeed X and S 2 , although they are functions only of the data and therefore show random variation, will converge to µ and σ 2 when we have a lot of data. Other distributions give different results. Take the exponential distribution 1 |x| f (x) = exp − 2a a where the expectation of |x| is the width parameter a. The pathological Cauchy distribution f (x) =
1 π(1 + x2 )
has the alarming property that the expectation of the average of N data is, again, the same Cauchy distribution; the location can apparently just
3.2 What should we expect of our statistics?
41
as well be estimated with one datum. The difficulty arises because the distribution has such wide wings. In astronomy, broad or even openended (power-law) distributions are common. It is worth checking any piece of remembered statistics, as it is almost certain to be based on the Gaussian distribution.
Other expectations of theoretical importance are known as the nth central moments: µn = (x − µ)n g(x) dx (3.7) where g is some probability distribution. They are estimated analogously by suitable averages to the way in which mean and variance were estimated in the previous example. They are sometimes useful for characterizing the shape of distributions, although they are very sensitive to outliers. Two descriptors using moments are common: skewness, β1 = µ23 , indicates deviation from symmetry (= 0 for symmetry about µ); and kurtosis, β2 = µ4 /µ32 , indicates degree of peakiness (= 3 for the Gaussian distribution). The Chebyshev inequality is sometimes useful: for any positive integer n, and data X drawn from a distribution of mean µ and variance σ 2 , prob[|X − µ| > nσ] ≤
1 . n
(3.8)
This is very conservative but is sometimes better than nothing as an estimate. 3.2 What should we expect of our statistics? We have but a few of the data Xi but we want to know how all of them are organized; we want their probability or frequency distribution and we want it for as little effort as we can get away with (efficiently) and as accurately as possible (robustly). Suppose, for instance, that we are drawing samples from a population obeying a Gaussian defined by µ = 0, σ = 1. Figure 3.1 conveys some indication of how the size of sample would affect estimates of these parameters. There are, then, at least four requirements for statistics. (i) They should be unbiased, meaning that the expectation value of the statistic turns out to be the true value. For the Gaussian
42
Statistics and expectations
Fig. 3.1. Xi drawn at random from a Gaussian distribution of σ = 1: (a) 20 values, (b) 100 values, (c) 500 values, (d) 2500 values. The average values of xi are 0.003, 0.080, −0.032 and −0.005; the median values 0.121, 0.058, −0.069 and −0.003; and the rms values 0.968, 1.017, 0.986, and 1.001. Solid curves represent Gaussians of unit area and standard deviation.
distribution (Section 2.4.2.3), for data Xi , X is indeed an unbiased estimate of the mean µ, but the unbiased estimate of the variance σ 2 is N 1 σs2 = (Xi − X)2 N − 1 i=1 which differs from the expectation value of S 2 by the factor N/(N − 1). The factor is confusing: σs2 , sometimes referred to as the sample variance, is the estimator for the population variance σ 2 . (The difference is understandable as follows. The Xi of our sample are first used to get X, an estimate of µ, and although this is an unbiased estimate of µ it is the estimate which yields a minimum value from the sum of the squares of the deviations of the sample, and thus a low estimate of the variance. The theory provides the appropriate correction factor N/(N − 1); of course the difference disappears as N → ∞.)
3.3 Simple error analysis
43
(ii) They should be consistent, the case if the descriptor for an arbitrarily large sample size gives the true answer. As we have seen, the rms is a consistent measure of the standard deviation of a Gaussian distribution in that it gives the right answer for large N ; but it is a biased estimator for small N unless modified by the factors just discussed. (iii) The statistic should obey closeness, yielding the smallest possible deviation from the truth. The Cauchy distribution (Section 3.1) looks innocent enough, somewhat similar to a Gaussian, even. But with infinite variance, trying to estimate dispersion via the standard deviation would yield massive scatter and little information. (iv) The statistic should be robust. For example, if we have a fundamentally symmetric distribution of data but a few experimental errors creep in, outliers appearing at the ends of the distribution, then as a measure of central location the median is far more robust than the average – it is less affected by the outliers. 3.3 Simple error analysis 3.3.1 Random or systematic? The average is a very common statistic; it is what we are doing all the time, for example, in ‘integrating’ on a faint object. The variance on the average is
2 N 1 2 Sm =E Xi − µ N i=1 which, after some manipulation, is σ2 1 2 Sm = E[(Xi − µ)(Xj − µ)]. + 2 N N
(3.9)
i=j
Neglecting the last term for the moment, the first term expresses generally held belief – the error on the mean of some data diminishes, √ like N , as the amount of data is increased. This is one of the most important tenets of observational astronomy. Now for the last term: apart from infinite√variances (e.g. the Cauchy distribution), the familiar and comforting N result holds only when this last term is zero. The term contains the covariance, defined as cov[Xi , Xj ] = E[(Xi − µi )(Xj − µj )];
(3.10)
44
Statistics and expectations
it is closely related to the correlation coefficient between xi and xj (Section 4.2). We are keeping the subscripts now because of the possibility that the data from the ith pixel, spectral channel, or time slot, are not independent of the data from the jth position. In the simplest cases, the data are independent and identically distributed (probability of Xi and Xj = probability of Xi × probability of Xj ) and then the covariance √ is zero. This is a condition (probably the likeliest) for the familiar N averaging away of noise; our assumption is that noise from one datum to the next or one pixel to the next is independent.
Suppose we had a time series, say of photometric measurements Xi . Here the i’s index time of observation. It might be a reasonable assumption that the measurements were identically distributed and independent of each other. In this case, the probability distribution would be the same for each time, and so can just be written g(x | parameters). The covariance term is then just
EXAMPLE
cov[Xi , Xj ] = E[(Xi − µi )(Xj − µj )] = (xi − µi )g(xi | . . .) dxi (xj − µj )g(xj | . . .) dxj =0
(3.11)
because, by definition of µ, each integral must separately be zero.
Often this simple situation does not apply. One possibility is that cov[xi , xj ] depends only on a ‘distance’ (i − j). If the data are indexed in some meaningful way, for example as a time series, the data are called stationary. As a second possibility, in photometric work it is quite likely that if one measurement is low, because of cloud, then the next few will be low too. (We speak of the dreaded 1/f noise; more of this in Section 8.8.) Then the probability distribution becomes multivariate and the simple factorizations do not apply: E[(Xi − µi )(Xj − µj )] = (xi − µi )(xj − µj )g(xi , xj | . . .) dxi dxj so that we need to know more about the observational errors – in other words, how to write down g(xi , xj | . . .) – before we can assess how the
3.3 Simple error analysis
45
average of the data will behave. In these more complicated cases, the √ averaging away is almost certain to be slower than N . A common distinction is made in experimental subjects between random and systematic errors, random errors being considered as those √ showing the N diminution. In reality there is a continuum, with the covariance frequently non-zero. At the other far extreme, systematic errors persist no matter how much data are collected. If you are observing Arcturus when you should be observing Vega, the errors will never average away no matter how persistent you are. Systematic errors can only be reduced by thorough understanding of the experimental equipment and circumstances; ‘random’ errors may be more or less random, depending on how correlated they are with each other.
3.3.2 Error propagation Often the thing we need to know is some more or less complicated function of the measured data. Knowing data error, how do we estimate error in the desired quantity? If the errors are small, by far the easiest way is to use a Taylor expansion. Suppose we measure variables x, y, z, . . . with independent errors δX, δY, δZ, . . . and we are interested in some function f (x, y, z, . . .). The change in f caused by the errors is, to first order, ∂f ∂f ∂f δF = δX + δY + δZ + · · · ∂x x=X ∂y y=Y ∂z z=Z The variance on a sum is the sum of the variances of the individual terms (because the errors are assumed to be independent) so we get 2 2 2 ∂f ∂f ∂f 2 2 var[f ] = σx + σy + σz2 + · · · ∂x ∂y ∂z x=X
y=Y
z=Z
(3.12) where the σ represent the variances in each of the variables. These considerations lead to a well-known result for combining measurements: if we have n independent estimates, say Xj , each having an associated error σj , the best combined estimate is the weighted mean, n j=1 wj X j X w = n j=1 wj where the weights are given by wj = 1/σj2 , the reciprocals of the sample
46
Statistics and expectations
variances. The best estimate of the variance of X w is 2 σw = n
1
j=1
EXAMPLES
1/σj2
.
Suppose (i) f (x, y) = x/y. Then the rule gives us imme-
diately var[f ] σx 2 = + f2 x
σy y
2 ;
we simply add up the relative errors in quadrature. If (ii) f (x) = log x then the rule gives σ 2 x var[f ] = x and the error in the log is just the relative error in the quantity we have measured.
3.3.3 Combining distributions Often this method is not good enough – we may need to know details of the probability distribution of the derived quantity. The simplest case is a transformation from the measured x, with probability distribution g, to some derived quantity f (x) with probability distribution h. Since probability is conserved, we have the requirement that h(f ) df = g(x) dx
(3.13)
so that h involves the derivative df /dx. Some care may be needed in applying this simple rule if the function f is not monotonic.
Suppose we are taking the logarithm of some exponentially distributed data. Here g(x) = exp(−x) for positive x, and f (x) = log(x). Applying our rule gives
EXAMPLE
h(f ) = exp(− exp(f )) exp(f ) which, as we might expect (Fig. 3.2) has a pronounced tail to negative values and is correctly normalized to unity. Our simpler methods would give us δh = δx/x, which evidently cannot give a good representation of the asymmetry of h. Quoting ‘h ± δh’ is clearly not very informative.
3.3 Simple error analysis
47
Probability density
0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 −4
−2
0 x
2
4
Fig. 3.2. The probability distribution of the logarithm of data drawn from an exponential distribution.
This technique rapidly becomes difficult to apply for more than one variable, but results for some useful cases are as follows. (1) Suppose we have two identically distributed independent variables x and y, both with distribution function g. What is the distribution of their sum z = x + y? For each x, we have to add up the probabilities of all the numbers y = z − x that yield the z we are interested in. The probability distribution h(z) is therefore h(z) = g(z − x)g(x) dx (3.14) where the probabilities are simply multiplied because of the assumption of independence. h is therefore the autocorrelation (Section 8.2) of g. The result generalizes to the sum of many variables, and is often best calculated with the aid of the Fourier transform (Section 8.2) of the distribution g. This transform is sometimes called the characteristic function. (2) Quite often we need the distribution of the product or quotient of two variables. Without details, the results are as follows. For z = xy, the distribution of z is 1 h(z) = g(x)g(z/x) dx (3.15) |x| and of z = x/y is
h(z) =
|x|g(x)g(zx) dx.
(3.16)
In almost any case of interest, these integrals are too hard to do analytically.
48
Statistics and expectations
One exception of interest is the product of two Gaussian variables of zero mean; this has applicability for a radio-astronomical correlator, for instance. Leaving out the mathematical details, the result emerges in the form of a modified Bessel function. The input Gaussians are of zero mean and variance σ 2 . The distribution of the product is 2 |z| h(z) = K 0 πσ 2 σ2
EXAMPLES
which as Fig. 3.3 shows is quite unlike a Gaussian. It has a logarithmic singularity at zero but is normalized to unity.
Probability density
1 0.8 0.6 0.4 0.2 0 −4
−2
0 x, z
2
4
Fig. 3.3. The probability distribution of the product of two identical Gaussians – the original Gaussian is the dashed curve.
The case of the ratio is equally instructive. Here we get h(z) =
1 1 , π 1 + z2
a Cauchy distribution. It has infinite variance and, as we see in Fig. 3.4, the variance of the original Gaussian surprisingly does not appear in the answer. This is a somewhat unrealistic case – it corresponds to forming the ratio of data of zero signal-to-noise ratio – but illustrates that ratios involving low signal-to-noise are likely to have very broad wings. The Bessel function distribution will, on average, succumb to the central limit theorem; this is not the case for the Cauchy distribution. In general, deviations from Normality will occur in the tails of distributions, the outliers that are so well known to all experimentalists.
3.4 Some statistics, and their distributions
49
Probability density
0.4 0.3 0.2 0.1 0 −4
−2
0 x, z
2
4
Fig. 3.4. The probability distribution of the ratio of two identical Gaussian variables the original Gaussian is the dashed curve.
3.4 Some statistics, and their distributions For N data Xi , some useful statistics are the average, the sample variance, and the order statistics. We have already met the first two; they acquire their importance because of their relationship to the parameters of the Gaussian. If the Xi are independent and identically distributed Gaussian variables, where the original Gaussian has mean µ and variance σ 2 , then: (i) The average X obeys a Gaussian distribution around µ, with variance σ 2 /N . We have met this result before (Section 2.5). (ii) The sample variance σs2 is distributed like σ 2 χ2 /(N − 1), where the chi-square variable has N −1 degrees of freedom (Table A2.6). (iii) The ratio √ N (X − µ) σs2 is distributed like the t statistic, with N − 1 degrees of freedom. This ratio has an obvious usefulness, telling us how far our average might be from the true mean (Table A2.3). (iv) If we have two samples (size N and M ) drawn from the same Gaussian distribution, then the ratio of the sample variances σs21 and σs22 follows an F distribution. This allows us to check if the data were indeed drawn from Gaussians of the same width (Section 5.2 and Table A2.4).
50
Statistics and expectations
The order statistics are simply the result of arranging the data Xi in order of size, relabelled as Y1 , Y2 , . . . So Y1 is the smallest value of X, and YN the largest. Maximum values are often of interest, and the median YN/2 (N even) is a useful robust indicator of location. We might also form robust estimates of widths by using order statistics to find the range containing, say, 50 per cent of the data. Both the density and the cumulative distribution are therefore of interest. Suppose the distribution of x is f (x), with cumulative distribution F (x). Then the distribution gn of the nth order statistic is gn (y) =
N! [F (y)]n−1 [1 − F (y)]N −n f (y) (n − 1)!(N − n)!
and the cumulative distribution is N N Gn (y) = [F (y)]j [1 − F (y)]N −j . j j=n
(3.17)
(3.18)
The Schechter luminosity function xγ exp(−x/x∗ ) is a useful model of the luminosity function for field galaxies. The observed value of γ is close to unity, but we will take γ = 1/2 for convenience in ensuring the distribution can be normalized over the range zero to infinity; we also take x∗ = 1. If we select 10 galaxies from this distribution, the maximum of the 10 will follow the distribution shown in Fig. 3.5. We
EXAMPLE
Probability density
1 0.8 0.6 0.4 0.2 0
1
2 3 Luminosity
4
5
Fig. 3.5. The Schechter luminosity function (solid curve) and the distribution of the maximum of 10 and 100 samples from the distribution, plotted as shortand long-dash curves respectively.
3.5 Uses of statistics
51
see that the distribution is quite different from the Schechter function, with a peak quite close to x∗ . If we choose 100 galaxies, then of course the distribution moves to brighter values.
3.5 Uses of statistics So far, we have concentrated on defining statistics and noticing that they (a) may estimate parameters of distributions and (b) will be distributed in some more or less complicated way themselves. Their use then parallels the Bayesian method. First, we may use them to estimate parameters; but the way in which they do this is more subtle than the Bayesian case. We do not get a probability distribution for the parameter of interest, but a distribution of the statistic, given the parameter. As noted in the introductory section of this chapter, the confidence interval is the usual way of making use of a statistic as an estimator. Second, we may test hypotheses. This again parallels the Bayesian case, but the methods are much further apart conceptually. Recall the case discussed in Section 2.5, where we have estimated parameters α and β. Using statistics, we would have two data combinations A (estimating α) and B (estimating β). How would we answer a question like ‘is α > β’ in this approach? The classical method entails finding some new combination, say t = A − B, and then computing its distribution on the hypothesis that α − β = 0. We then find the probability of the observed value of t, or bigger, occurring on this hypothesis; and if the probability is small, we would conclude that the data were unlikely to have occurred by chance. The hint, of course, is that indeed α > β, but we do not know the probability of this. This classical approach is the basis of numerous useful tests, and we discuss some of them in detail in later Chapters 4 and 5. However, there is no doubt that the method does not quite seem to answer the question we had in mind, although often its results are indistinguishable from the more intelligible Bayesian approach. The same decisions get taken. Perhaps the most difficult part of this testing procedure is the implicit use of data corresponding to events that did not occur – the ‘observed value of t, or bigger’ referred to. Jeffreys (1961) wrote ‘. . . a
52
Statistics and expectations
hypothesis that may be true may be rejected because it has predicted observable results that have not occurred. This seems a remarkable procedure.’ However, using large but unobserved values of the test statistic usually does not matter much; in cases of interest, our statistic will be unlikely anyway, and larger values will be even less likely.
Exercises 3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
Means and variances. Find the mean and variance of a Poisson distribution and of a power law; find the variance (= ∞) of a Cauchy distribution. Simple error analysis. Derive the well-known results for error combining, for two products, and the sum and difference of two quantities, from the Taylor expansion of Section 3.3.2. Combining Gaussian variables. Use the result of Section 3.3.2 for errors on z when z = x + y to find the distribution of the sum of two Gaussian variables. Average of Cauchy variables. Show that the average value of Cauchy-distributed variables has the same distribution as the original data. Use characteristic functions and the convolution theorem. Find a better location estimator. Poisson statistics. Draw random numbers from Poisson distributions (Section 6.5) with µ = 10 and µ = 100. Take 10 or 100 samples, find the average and the rms scatter. How close is √ the scatter to average? Robust statistics. Make a Gaussian with outliers by combining two Gaussians, one of unit variance, one three times wider. Leave the relative weight of the wide Gaussian as a parameter. Compare the mean deviation with the rms, for various relative weights. How sensitive are the two measures of scatter to outliers? Repeat the exercise, with a width derived from order statistics. Change of variable. Suppose that φ is uniformly distributed between zero and 2π. Find the distribution of sin φ. How could you find the distribution of a sum of sines of independent random angles? Order statistics. We record a burst of N neutrinos from a supernova, and the probability of recording a neutrino at time t
Exercises
53
is, in suitable units, exp(t − t0 ) where t0 is the time of emission. The maximum-likelihood estimate of t0 is just T1 , the time of arrival of the first neutrino. Use order statistics (Section 3.4) to show that the average value of T1 is just t0 + N1 . Is this MLE biased, but consistent (i.e. the correct answer as N → ∞)?
4 Correlation and association
Arguing that the trial judge had failed to explain clearly the use of Bayes’ theorem, the defence lodged an appeal. But in a bizarre irony, the Appeal Court last month upheld the appeal and ordered a retrial – on the grounds that the original judge had spent too much time explaining the scientific assessment of evidence. In their ruling, the Appeal judges said: ‘To introduce Bayes’ theorem, or any similar method, into a criminal trial plunges the jury into inappropriate and unnecessary realms of theory and complexity’. (Robert Matthews, New Scientist 1996)
When we make a set of measurements, it is instinct to try to correlate the observations with other results. One or more motives may be involved in this instinct: for instance we might wish (1) to check that other observers’ measurements are reasonable, (2) to check that our measurements are reasonable, (3) to test a hypothesis, perhaps one for which the observations were explicitly made, or (4) in the absence of any hypothesis, any knowledge, or anything better to do with the data, to find if they are correlated with other results in the hope of discovering some new and universal truth.
4.1 The fishing trip Take the last point first. Suppose that we have plotted something against something, on a fishing expedition of this type. There are grave dangers on this expedition, and we must ask ourselves the following questions. (1) Does the eye see much correlation? If not, calculation of a formal correlation statistic is probably a waste of time. 54
4.1 The fishing trip
55
Fig. 4.1. Radio luminosities of 3CR radio sources versus distance modulus. The curved line represents the survey limit, the limit imposed by forming a catalogue from a flux-limited sample (Section 7.2).
(2) Could the apparent correlation be due to selection effects? Consider for instance the beautiful correlation in Fig. 4.1, in which Sandage (1972) plotted radio luminosities of sources in the 3CR catalogue as a function of distance modulus. At first sight, it proves luminosity evolution for radio sources. Are the more distant objects (at earlier epochs) clearly not the more powerful? In fact, as Sandage recognized, it proves nothing of the kind. The sample is flux- (or apparent intensity) limited; the solid line shows the flux-density limit of the 3CR catalogue. The lower right-hand region can never be populated; such objects are too faint to show above the limit of the 3CR catalogue. But what about the upper left? Provided that the luminosity function (the true space density in objects per megaparsec3 ) slopes downward with increasing luminosity, the objects are bound to crowd towards the line. This is about all that can be gleaned immediately from the diagram – the space density of powerful radio sources is less than the space density of their weaker brethren. Astronomers produce many plots of this type, and will describe purported correlations in terms such as ‘The lower right-hand region of the diagram is unpopulated because of the detection limit, but there is no reason why objects in the upper left-hand region should have escaped
56
Correlation and association
Fig. 4.2. Dodgy correlations: in each case formal calculation will indicate that a correlation exists to a high degree of significance.
detection’. True, but nor can they escape probability; the upper left of Sandage’s diagram is not filled with QSOs and radio galaxies because we need to sample large spheres about us to have a hope of encountering a powerful radio source. Small spheres, corresponding to small redshifts and distance moduli, will yield only low-luminosity radio sources because their space density is so much the higher. The lesson applies to any proposed correlation for variables with steep probability density functions dependent upon one of the variables plotted. (3) If we are happy about (2), we can try formal calculation of the significance of the correlation as described in Section 4.2. Further, if there is a correlation, does the regression line (Section 6.2) make sense? (4) If we are still happy, we must return to the plot to ask if the formal result is realistic. A rule of thumb: if 10 per cent of the points are grouped by themselves so that covering them with the thumb destroys the correlation to the eye, then we should doubt it, no matter what significance level we have found. Beware in particular of plots which look like those of Fig. 4.2, plots which strongly suggest selection effects, data errors, or some other form of statistical conspiracy. (5) If we are still confident, we must remember that a correlation does not prove a causal connection. The essential point is that correlation may simply indicate a dependence of both variables on a third variable. Cigarette manufacturers said so for years; but finding the physical attribute which caused heart/lung disease and the desire to smoke proved difficult. But there are many famous instances, e.g. the correlation
4.2 Testing for correlation
57
between quality of children’s handwriting and their height, and between the size of feet in China and the price of fish in Billingsgate Market. For the former the hidden variable is age (Are tall children cleverer? No, but older), while for the latter it is time. There are in fact ways of searching for intrinsic correlation between variables when they are known to depend mutually upon a third variable. The problem, however, when on the fishing trip, is how to know about a third variable, how to identify it when we might suspect that it is lurking. We consider it further in Sections 4.3 and 4.5. Finally we must not get too discouraged by all the foregoing. Consider Fig. 4.3, a ragged correlation if ever there was one, although there are no nasty groupings of the type rejected by the rule of thumb. It is in fact one of the earliest ‘Hubble diagrams’ – the discovery of the recession of the nebulae, and the expanding universe (Hubble 1936).
Fig. 4.3. (a) An early Hubble diagram (Hubble 1936); recession velocities of a sample of 24 galaxies versus distance measure. (b) The same plot but with data normalized by standard deviation; the lines represent principal components, as described in Section 4.5.
4.2 Testing for correlation In dealing with correlations we encounter in detail many important aspects of the use of probability and statistics. The foregoing problem appears simple: we have a set of N measurements (Xi , Yi ) and we ask (formally) if they are related to each other. To make progress we have to make ‘related’ more precise. The bestdeveloped way of doing this – although not necessarily relevant – is to model our data as a bivariate or joint Gaussian of correlation
58
Correlation and association
coefficient ρ: prob(x, y | σx , σy , ρ) = × exp
2πσx σy
1
1 − ρ2
(4.1)
−1 (y − µy )2 2ρxy (x − µx )2 . + − 2(1 − ρ2 ) σx2 σy2 σx σy
This model is so well developed that ‘correlation’ and ‘ρ = 0’ are nearly synonymous; if ρ → 0 there is little correlation, while if ρ → 1 the correlation is perfect; see Fig. 4.4.
Fig. 4.4. Linear contours of the bivariate Gaussian probability distribution; the near-circular contours represent ρ = 0.01, a bivariate distribution with little connection between x and y, while the highly elliptical contours represent ρ = 0.99, indicative of a strong correlation between x and y. Negative values of ρ reverse the tilt, and indicate what is loosely referred to as anticorrelation.
The parameter ρ is the correlation coefficient, and in the above formulation, it is given by ρ=
cov[x, y] σx σy
(4.2)
where cov is the covariance (Section 3.3.1) of x and y, and σx2 and σy2
4.2 Testing for correlation
59
are the variances. The correlation coefficient can be estimated by N (Xi − X)(Yi − Y ) r = i=1 . (4.3) N N 2 2 (X − X) (Y − Y ) i i i=1 i=1 r is known as the Pearson product-moment correlation coefficient (Fisher 1944). The contours of Fig. 4.4 will have dropped by 1/e from the maximum at the origin when 2 1 x y2 2ρxy = 1, (4.4) + 2− 1 − ρ2 σx2 σy σx σy or in matrix notation, when 1 (x y) 1 − ρ2
1 2 σx − σxρσy
− σxρσy 1 σy2
x = 1. y
(4.5)
The inverse of the central matrix is known as the covariance matrix or error matrix σx2 cov[x, y] C= . (4.6) cov[x, y] σy2 The off-diagonal elements of the covariance matrix can be estimated by 1 (Xi − Xi )(Xj − Xj ). N −1 The matrix is particularly valuable in calculating propagation of errors, but there are numerous applications, for example in principal component analysis (Section 4.5) and in maximum-likelihood modelling (Section 6.1). The multivariate Gaussian is one example of a class of multivariate distribution functions that depend only on the data vector x via a socalled quadratic form xT C x. The multivariate Gaussian is the most familiar of these. To return to the point at issue: what we really want to know is whether or not ρ = 0; it is this condition for which we are testing. Using the bivariate Gaussian is a very specific model; a Gaussian is assumed, it allows only two variances, and assumes that both x and y are random variables. Thus σx and σy include both the errors in the data, and their
60
Correlation and association
intrinsic scatter – all presumed Gaussian. The model does not apply, for example, to data where the x-values are well defined and there are ‘errors’ only in y, perhaps different at different x. In such cases we would use model fitting, perhaps of a straight line (Sections 6.1 and 6.2). This is a different problem. These effects mean that we have to approach the correlation coefficient with caution, as the way we set up our experiment may result in graphs like those of Figs. 4.1 or 4.2. As always, there are two quite different ways of proceeding from this point, Bayesian and non-Bayesian.
4.2.1 Bayesian correlation testing The Bayesian approach is to use Bayes’ theorem to extract the probability distribution for ρ from the likelihood of the data and suitable priors. Since we want to know about ρ independently of any inference about the means and variances, we have to integrate these ‘nuisance variables’ out of the full posterior probability prob(ρ, σx , σy , µx , µy | data). For the bivariate Gaussian model, the result is given by Jeffreys (1961) as (1 − ρ2 )(N −1)/2 1 + rρ 1 prob(ρ | data) ∝ + . . . . (4.7) 1+ n − 1/2 8 (1 − ρr)N −3/2 The Bayesian test for correlation is thus simple: compute r from the (Xi , Yi ), and calculate prob(ρ) for the range of ρ of interest.
We generated 50 samples from a bivariate t distribution with three degrees of freedom. The true correlation coefficient was 0.5. The large tails of the distribution produce outliers, not accounted for by the assumed Gaussian used in interpreting the r statistic. Figure 4.5 shows what equation (4.7) gives: the distribution of ρ peaks at around 0.2. If now we remove the samples outside 4σ, the distribution peaks at around 0.5 and is appreciably narrower. The method is thus fairly robust, although obviously affected by being used with the ‘wrong’ distribution.
EXAMPLE
Given this probability distribution for ρ, we can answer questions like ‘what is the probability that ρ > 0.5?’ or (perhaps more usefully) ‘what is the probability that ρ from dataset A is bigger than ρ from dataset B?’ (see Section 2.5). As is often the case, the utility of the Bayesian
4.2 Testing for correlation
61
Fig. 4.5. Fifty Xi , Yi chosen at random from a bivariate Gaussian with ρ = 0.5, with some outliers added. The Jeffreys probability distribution of the correlation coefficient ρ is shown, peaking at around 0.2 for the upper panel. The data have been restricted to ±4σ in the lower panel; the distribution now peaks at 0.44.
approach is not that prior information is accurately incorporated, but rather that we get an answer to the question we really want to ask. Jeffreys used a uniform prior for ρ – not obviously justifiable, and certainly not correct if ρ is close to 1 or −1, as he points out. But in these cases a statistical test is a waste of time anyway.
An interesting use of Jeffreys’s distribution is to calculate the probability that ρ is positive, as a function of sample size (Fig. 4.6.) This tells us how much data we need to be confident of detecting correlations.
EXAMPLE
4.2.2 The classical approach to correlation testing The alternative approach to the correlation problem starts by regarding ρ as a fixed quantity, not a variable about which probabilistic statements
62
Correlation and association 1 0.98 0.96 0.94 0.92 10
15 20 30 Number of observations
50
Fig. 4.6. The probability of ρ being positive, as a function of sample size, for r-values of 0.25 (lowest curve), 0.5 and 0.75 (uppermost curve).
might be made. This approach therefore arrives at the probability of the data, given ρ (and of course the background hypothesis that a bivariate Gaussian is adequate). The result (Fisher 1944) is prob(r | ρ, H) ∝
(1 − ρ2 )(N −1)/2 (1 − r2 )(N −4)/2 (1 − ρr)N −3/2 1 1 + rρ × 1+ + ··· . N − 1/2 8
(4.8)
What can we do with this answer? The standard approach is to pick the easy ‘null hypothesis’ ρ = 0, compute r, and then compute the probability, under the null hypothesis, of r being this big or bigger. If this probability is very small, we may feel that the null hypothesis is rather unlikely. The standard parametric test is to attempt to reject the hypothesis that ρ = 0 and we do this by computing r. The standard deviation in r is (1 − r2 ) σr = √ . N −1
(4.9)
Note that −1 < r < 1; r = 0 for no correlation. To test the significance of a non-zero value for r, compute √ r N −2 t= √ (4.10) 1 − r2 which obeys the probability distribution of the ‘Student’s’ t statistic1 1
After its discoverer W. S. Gosset (1876–1937), who developed the test while working on quality control sampling for Guinness. For reasons of industrial secrecy,
4.2 Testing for correlation
63
with N − 2 degrees of freedom. (The transformation simply allows us to use tables of t.) We are hypothesis testing now, and the methodology is described more systematically in Section 4.1. Consult Table A2.3, the table of critical values for t; if t exceeds that corresponding to a critical value of the probability (two-tailed test), then the hypothesis that the variables are unrelated can be rejected at the specified level of significance. This level of significance (say 1 per cent, or 5 per cent) is the maximum probability which we are willing to risk in deciding to reject the null hypothesis (no correlation) when it is in fact true. This approach has probably not answered the question – we embark on this sort of investigation when it is apparent that the data contain correlations; we merely want some justification by knowing ‘how much’. Also, the inclusion in the testing procedure of values of r that have not been observed poses the usual difficulties. The test is widely used, and is formally powerful. But as one statistics book says ‘There are data to which this kind of correlation method cannot be applied.’ This is a gross understatement. The data must be on continuous scales, obviously. The relation between them must be linear. (How would we know this? In many cases in astronomy we change the scales at will (log–log, log–linear, etc.) to give a roughly linear appearance to our plots.) The data must be drawn from Normally distributed populations. (How would we know this? Certainly if we have changed our data axes to log form, there must be doubt.) They must be free from restrictions in variability or groupings. There are parametric tests that help: the F test for non-linearity and the correlation ratio test which gets around non-linearity. However, to circumvent the problems it is far better to go to a non-parametric test. These permit additional tests on data which are not numerically defined (binned data, or ranked data), so that in some instances they may be the only alternative.
4.2.3 Correlation testing: classical, non-parametric The best-known non-parametric test consists of computing the Spearman rank correlation coefficient (Conover 1999; Siegel & Castellan 1988): N (Xi − Yi )2 rs = 1 − 6 N3 − N
(4.11)
Gosset was required to publish under a pseudonym; he chose ‘Student’, which he used for years in correspondence with his (former) professor at Oxford, Karl Pearson.
64
Correlation and association
where there are N data pairs, and the N values of each of the two variables are ranked so that (Xi , Yi ) represents the ranks of the variables for the ith pair, 1 < Xi < N , 1 < Yi < N . The range is 0 < rs < 1; a high value indicates significant correlation. To find how significant, refer the computed rs to Table A2.5, a table of critical values of rs applicable for 4 ≤ N ≤ 30. If rs exceeds an appropriate critical value, the hypothesis that the variables are unrelated is rejected at that level of significance. If N exceeds 30, compute N −2 tr = r s , (4.12) 1 − rs2 a statistic whose distribution for large N asymptotically approaches that of the t statistic with N −2 degrees of freedom. The significance of tr may be found from Table A2.3, and this represents the associated probability under the hypothesis that the variables are unrelated. How does use of rs compare with use of r, the most powerful parametric test for correlation? Very well: the efficiency is 91 per cent. This means that if we apply rs to a population for which we have a data pair (xi , yi ) for each object and both variables are Normally distributed, we will need on average 100 (xi , yi ) for rs to reveal that correlation at the same level of significance which r attains for 91 (xi , yi ) pairs. The moral is that if in doubt, little is lost by going for the non-parametric test. The Kendall rank correlation coefficient does the same thing as rs , and with the same efficiency (Siegel & Castellan 1988).
A ‘correlation’ at the notorious 2σ level is shown in Fig. 4.7. Here, rs = 0.28, N = 55, and the hypothesis that the variables are unrelated is rejected at the 5 per cent level of significance. Here we have no idea of the underlying distributions; nor are we clear about the nature of the axes. The assumption of a bivariate Gaussian distribution would be rash in the extreme, especially in view of a uniformly filled Universe producing a V /Vmax statistic uniformly distributed between 0 and 1 (Schmidt 1968). The Vmax method is discussed in Section 7.3.
EXAMPLE
There is yet another way, the permutation test. In the case of correlation analysis, we have data (X1 , Y1 ), (X2 , Y2 ), . . . and we wish to test the null hypothesis that x and y are uncorrelated. In this regard if we
4.2 Testing for correlation
65
Fig. 4.7. V /Vmax as a function of high-frequency spectral index for a sample of radio quasars selected from the Parkes 2.7-GHz survey.
have some home-made test statistic η, we can calculate its distribution, on the assumption of the null hypothesis, by simply calculating its value for many permutations of the x’s amongst the y’s. For any reasonable dataset there will be far more possible permutations than we can reasonably explore, but choosing a random set will give an adequate estimate of the distribution of the test statistic. If it turns out that the observed value of η is very improbable, under the null hypothesis, we may be interested in estimating the distribution for non-zero correlation. This is a route to useful Bayesian analysis, of the kind we described for the correlation coefficient ρ. Here Monte Carlo simulation (Section 6.5) will come into its own, allowing us to explore a wide range of parameter space, so building up the posterior distribution prob(parameters/η). These methods can be used to derive distributions of statistics such as Spearman’s or Kendall’s correlation coefficients in cases when a correlation is apparently present.
4.2.4 Correlation testing: Bayesian versus non-Bayesian tests Let us be clear: the non-parametric tests circumvent some of the issues involved in the non-Bayesian approach, but they have no bearing on the
66
Correlation and association
fundamental issue – what was the real question? However, the Bayesian approach, strong in answering the real question, forces reliance on a model. There is rather little difference, in practice, between the Fisher test and results from Jeffreys’s distribution. We can show this with some random Gaussian data with a correlation of zero. In the standard way, we can use the r distribution to find the probability of r being as large, or larger, than we observe, on the hypothesis that ρ = 0. If this probability is small, the test is hinting at the possibility that the correlation is actually positive. Therefore we compare with the probability, from the Jeffreys distribution, that ρ is positive. If the probability from Fisher’s r distribution is small we expect the probability from ρ to be large; and in fact we can see, either from simulations or from the algebraic form of the distributions, that the sum of these two probabilities is close to 1. In other words, interpreting the standard Fisher test (illegally!) to be telling us the chance that ρ is positive, actually works very well.
4.3 Partial correlation The ‘lurking third variable’ can be dealt with (provided that its influence is recognized in the first place) by partial correlation, in which the ‘partial’ correlation between two variables is considered by nullifying the effects of the third (or fourth, or more) variable upon the variables being considered. Partial correlation is a science in itself; it is covered in both parametric and non-parametric forms by Stuart & Ord (1994), Macklin (1982), and Siegel & Castellan (1988). In the parametric form, consider a sample of N objects for which parameters x1 , x2 , and x3 have been measured. The first-order partial correlation coefficient between variables x1 and x2 is r12.3 =
r12 − r13 r23 2 )(1 − r 2 ) (1 − r13 23
(4.13)
where the r are the product-moment coefficients defined in Section 4.2.2. If there are four variables, then the second-order partial correlation coefficient is r12.3 − r14.3 r24.3 r12.34 = (4.14) 2 2 (1 − r14.3 )(1 − r24.3 ) where the correlation is being examined between x1 and x2 with x3
4.4 But what next?
67
and x4 held constant. Examination of the correlation between the other variables requires manipulation of the subscripts in the foregoing. And so forth for higher-order partial correlations between more than four variables, with the standard error of the partial correlation coefficients being given by σr12.34...m =
2 1 − r12.34...m √ N −m
where m is the number of variables involved. The significance then comes from the ‘Student’s’ t test as above.
Consider data from a sample of lads aged 12–19. The correlation between height and weight will be high because the older boys are taller on average. But with age held constant, the correlation would still be significantly positive because at all ages, taller boys tend to be heavier. In such a sample of 10, the correlation between height and weight (r12 ) is calculated as 0.78; between height and age (r13 ), 0.52, and between weight and age, r23 = 0.54. The first-order partial coefficient of correlation (Equation 4.13) is thus r12,3 = 0.69; σr12.3 = 0.198; and the correlation is significant at the level of 0.2 per cent. Consider further a measure of strength for each lad. The correlation between strength and height (r41 ) is 0.58; between strength and weight (r42 ) 0.72. Will lads of the same weight show a dependence of strength upon height? The answer is given by r41.2 = 0.042; the correlation between strength and height essentially vanishes and we would conclude that height as such has no bearing on strength; only by virtue of its correlation with weight does it show any correlation at all. As for second-order partials, is there a correlation between strength and age if height and weight are held constant? The raw correlation between age and strength was 0.29; the second-order partial also yields 0.29. It seemingly makes little difference if height and weight are allowed to vary; the relation between age and strength is the same.
EXAMPLE
4.4 But what next? If we have demonstrated a correlation, it is logical to ask what the correlation is, i.e. what is the law which relates the variables. It is
68
Correlation and association
common practice to dash off and fit a regression2 line, usually applying the method of least squares (Section 6.2). It is essential to note that this is model fitting now; the distinction between data modelling (Chapter 6) and hypothesis testing (here; and Chapter 5) is important. Before doing so, there are several considerations, most of which are addressed in more detail in Section 6.2. Are there better quantities to minimize than the squares of deviations? What errors result on the regression-line parameters? Why should the relation be linear? And – most crucial of all – what are we trying to find out? If we have found a correlation between x and y, which variable is dependent; do we want to know x on y or y on x? The coefficients are generally completely different. As an argument against blind application of correlation testing and line fitting, consider the famous Anscombe (1973) quartet, shown in Fig. 4.8. Anscombe’s point is the essential role of graphs in good statistical analysis. However, the examples illustrate other matters: the rule of thumb (Section 4.1), and the distinction between independence of data points and correlation. In more than one of Anscombe’s datasets the points are clearly related. They are far from independent, while not showing a particularly strong (formal) correlation. The upper right example in Figure 4.8 is a case in which a linear fit is of indifferent quality, while the choice of the ‘right’ relation between X and Y would result in a perfect fit. The quartet further emphasizes how dependent our analyses are on the assumption of Gaussianity: the covariance matrix, which intuitively we might expect to reflect some of the structure in the individual plots, is identical for each. Note that X independent of Y means prob(X, Y ) = prob(X)prob(Y ), or prob(X | Y ) = prob(X); while X correlated with Y means prob(X, Y ) = prob(X) prob(Y ) in a particular way, giving r = 0. It is perfectly possible to have prob(X, Y ) = prob(X)prob(Y ) and r = 0, the standard example being points distributed so as to form the Union Jack. If we simply wish to map the dependence of variables on each other with minimal judgemental input, it strongly suggested, here and in 2
Galton (1889) introduced the term regression; it is from his examination of the inheritance of stature. He found that the sons of fathers who deviate x inches from the mean height of all fathers themselves deviate from the mean height of all sons by less than x inches. There is what Galton termed a ‘regression to mediocrity’. The mathematicians who took up his challenge to analyse the correlation propagated his mediocre term, and we’re stuck with it.
4.5 Principal component analysis
69
Fig. 4.8. Anscombe’s quartet: four fictitious sets of 11 (Xi , Yi ), each with the same (X, Y ), identical coefficients of regression, regression lines, residuals in Y and estimated standard errors in slopes.
Section 6.2, that principal component analysis is the appropriate technique.
4.5 Principal component analysis Principal component analysis (PCA) is the ultimate correlation searcher when many variables are present. Given a sample of N objects with n parameters measured for each of them, how do we find what is correlated with what? What variables produce primary correlations, and what produce secondary, via the lurking third (or indeed n − 2) variables? PCA is one of a family of algorithms (known as multivariate statistics; see e.g. Manly (1994), Kendall (1980), Joliffe (2002)) designed for this situation. Its task is the following: given a sample of N objects with n measured variables xn for each, find a new set of ξn variables that are orthogonal (independent), each one a linear combination of the original variables: n
ξi = aij xj (4.15) j=1
70
Correlation and association
with values of aij such that the smallest number of new variables accounts for as much of the variance as possible. The ξi are the principal components. If most of the variance involves just a few of the n new variables, we have found a simplified description of the data. Finding which of the variables correlate (and how) may lead to that successful fishing expedition – we may have caught new physical insight. PCA may be described algebraically, through covariance matrices (Section 4.2), or geometrically. Taking the latter approach, consider the N objects represented by a large cloud in n-dimensional space. If two of the n parameters are correlated, the cloud is elongated along some direction in this space. PCA identifies these extension directions and uses them as a sequential set of axes, sequential in the sense that the most extended direction is identified first by minimizing the sums of squares of deviations. This direction forms the first principal component (or eigenvector 1), accounting for the largest single linear variation amongst the object properties. Then the (n − 1)-dimensional hyperplane orthogonal to the first principal component is considered and searched for the direction representing the greatest variance in (n − 1)-space; and so forth, defining a total of n orthogonal directions.
As an elementary PCA example via geometry, let us return to the early Hubble diagram of Fig. 4.3, 24 galaxies with two measured variables, velocity of recession v and distance d. It is standard practice to normalize by subtracting the means from each variable and to divide by the standard deviation, i.e. to plot vi = (vi − < v >)/σv versus di = (di − < d >)/σd , as shown in Fig. 4.3(b). Then we find the first principal component by simply rotating the axis through the origin to align with maximum elongation, the direction of apparent correlation, and we do this with least squares (Section 6.2) – maximizing the variance along PC1 is equivalent to minimizing the sums of the squares of the distances of the points from this line through the origin. The distance of a point from the direction PC1 (shown dotted in Fig. 4.3b) represents the value (score) of PC1 for that point. PC1 is clearly a linear combination of the two original variables; in fact it is v = d . Because the new coordinate system was found by simple rotation, distances from the origin are unchanged; the total variance of v and d is unchanged and is 2.0. The variance of PC1, the normalized distances squared from PC2, is 1.837. The remaining variance of the sample must be accounted for by
EXAMPLE
4.5 Principal component analysis
71
Table 4.1. Principal components from Fig. 4.3 PC1
PC2
Eigenvalue Proportion Cumulative
1.837 0.918 0.918
0.163 0.082 1.000
Variable
PC1
PC2
d (Mpc) v (km s−1 )
1.0 1.0
1.0 −1.0
the projection of data points onto the axis PC1, perpendicular to PC2; the length of these projections are the object’s values or scores of the second principal component, and this is verified as 0.163, with the sum of these variances 2.0 as expected. Table 4.1 sets out the results in the standard way of PCA.
Now consider the matrix approach. In the process of PCA the usual methodology is to construct the error matrix (Section 4.2), e.g. for the 2 two-variable case of the example, a(1, 1) = d , a(1, 2) = a(2, 1) = 2 v d , a(2, 2) = v . We then seek a principal axis transformation that makes the cross-terms vanish; we seek an axis transformation to rotate the ellipses of Fig. 4.4 so that the axes of the ellipses coincide with the principal axes of the coordinate system. This of course is simply done in matrix notation. We determine the eigenvalues of the error matrix and form its eigenvectors (readily shown for the example to be v = d and v = −d as seen in Fig. 4.3b). These eigenvectors then form the transpose matrix T , for variable transformation and axis rotation. The axis rotation diagonalizes the matrix, i.e. in the new axis system, the cross-terms are zero; we have rotated the axes until there is no x, y covariance. Note that for the purpose our set of data has been reduced from 48 numbers for the 24 galaxies to four numbers, a 2 × 2 matrix. How did this happen? PCA assumes that the covariance (or error) matrix suffices to describe the data; this is the case if the data are drawn from a multivariate Gaussian (Section 4.2, Fig. 4.4), or in general when a simple quadratic form, using the covariance matrix, can describe the distribution of the data. It is far from generally true that the clouds of
72
Correlation and association
points in most n-variate hyperspaces will be so simply distributed – see the following example. The distribution need not be symmetrical, for example. In multivariate datasets, the disparate units are taken care of by normalizing as in the above example: subtracting mean values and dividing by variances. This is not a prescription, however. For example, the variance for any particular variable might be dominated by a monstrous outlier which there are good grounds to reject. The choice of weights does therefore depend on familiarity with the data and preferences – there is plenty of room for subjectivity. It should also be noted that PCA is a linear analysis and tests need to be performed on the linearity of the principal components. For example, plotting the scores of PC1 versus PC2 should show a roughly Gaussian distribution consistent with ρ = 0. It may be apparent how to reject outliers or to transform coordinates to reduce the problem to a linear analysis. In large datasets such processes can reveal unusual objects.
Some PCA problems have a larger number of variables than input observables, p > n, resulting in singular matrices requiring modifications to standard techniques to solve the eigenvector equations (Wilkinson 1978; Mittaz, Penston & Snijders 1990). This situation occurs in spectral PCA for which the p variables are fluxes in p wavelength or frequency bins (Francis et al. 1992; Wills et al. 1997). The technique is ideal for dealing with a huge sample and was therefore adopted in the 2dF survey which aims to measure 250 000 galaxy spectra to provide a detailed picture of the galaxy distribution out to a redshift of 0.25. The PCA approach to 2dF galaxy classification is discussed in detail by Folkes et al. (1999). Figure 4.9, drawn from this paper, shows examples of 2dF spectra prepared for PCA, the mean spectrum, and the first three principal components. These three components represent the eigenvectors of the covariance matrix of these prepared spectra. In this example, the first PC accounts for 49.6 per cent of the variance; the first three components account for 65.8 per cent of the variance. Much of the remainder is due to noise. The key aspect Folkes et al. (1999) wished to address was how the luminosity function depends on galaxy type. The objects in the PC1– PC2 plane form a single cluster (Fig. 4.9, blue emission-line objects to the left, red objects with absorption lines to the right, and strong
EXAMPLE
4.5 Principal component analysis
73
0.4 0.2 0
0 Photon counts [Arbitrary units]
-0.1 -0.2 -0.3 -0.4 0.15 0.1 0.05 0 -0.05 1.6 1.4 1.2 1 0.8 0.6 0.4 4000
5000
6000
Wavelength
Fig. 4.9. Top left: examples of 2dF spectra prepared for PCA. Instrumental and atmospheric features have been removed, with the spectra transformed to the rest frame, resampled to 4˚ A bins and normalized to unit mean flux. Top right: the mean spectrum and first three principal components; the sign of the PCs is arbitrary. Below: distribution of 2dF galaxy spectra in the PC1–PC2 plane. Slanted lines divide the plane into the five spectral classes adopted by Folkes et al.
emission-line objects straggling downward). Five spectral classes were then adopted, shown by the slanted lines in this figure. Confirmation that these spectral classes correspond to morphological classification came from placing the 55 Kennicut (1992) standard galaxies into this plot; the five classes are roughly E/SO, Sa, Sb, Scd and Irr. The way ahead to use the PCA classes to work out luminosity functions for each is clear, and the punch line is that significantly different Schecter functions emerged for each class.
74
Correlation and association
Note how asymmetrical the distribution looks. This need not invalidate the analysis – here primarily one of classification – but the effectiveness must in general be reduced. Asymmetrical shapes in the PC planes must result in unquantifiable errors in the classification.
In addition to spectral classification and analysis, spectral time variability is amenable to PCA (Mittaz, Penston & Snijders 1990; Turler & Courvoisier 1998). Of course we would like to know if the PCs are ‘real’ and so some indication of the distribution of each one would be useful. This can be computed by a bootstrap (Section 6.6) on the original dataset. This will show how stable the eigenvectors and eigenvalues actually are, in particular whether the largest eigenvector is reliably detected. Exercises In the exercises denoted by (D), datasets are provided on the book’s website; or create your own. 4.1
4.2
4.3
Correlation testing (D). Consider the Hubble plot of Fig. 4.3. What is (a) the most likely value for ρ via the Jeffreys test, (b) the significance of the correlation via the standard Fisher test and (c) the significance via the Spearman rank test? Estimate distributions for these statistics with a bootstrap (Section 6.6), and compare the results with the standard tests. Permutation tests (D). (a) Take a small set of uncorrelated pairs (X, Y ), preferably non-Gaussian. By permutation methods on the computer, derive distributions of Fisher r, Spearman’s and Kendall’s statistics. (b) Try the same numerical experiment with correlated data, using the bootstrap and the jackknife to estimate distributions (Sections 6.6). Correlated non-Gaussian data are provided for the multivariate t distribution, which is Cauchy-like for one degree of freedom and becomes more Gaussian for larger degrees of freedom. How robust are the conclusions against outliers? Principal component analysis (D). Carry through a PCA on the data of the quasar sample given in Francis & Wills (1999). Compute errors with a bootstrap analysis or jackknife (Section 6.6).
Exercises 4.4
75
Lurking third variables. Consider the following correlations, and speculate on how a third variable might be involved. (a) During the Second World War, J.W. Tukey discovered a strong positive correlation between accuracy of high-altitude bombing and the presence of enemy fighter planes. (b) There is a wellknown correlation between stock market indices and the sunspot cycle. (c) The apparent angular size of radio sources shows a strong inverse correlation with radio luminosity.
5 Hypothesis testing
How do our data look? I’ve carried out a Kolmogorov–Smirnov test . . . Ah. That bad. (interchange between Peter Scheuer and his then student, CRJ)
It is often the case that we need to do sample comparison: we have someone else’s data to compare with ours; or someone else’s model to compare with our data; or even our data to compare with our model. We need to make the comparison and to decide something. We are doing hypothesis testing – are our data consistent with a model, with somebody else’s data? In searching for correlations as we were in Chapter 4, we were hypothesis testing; in the model fitting of Chapter 6 we are involved in data modelling and parameter estimation. Classical methods of hypothesis testing may be either parametric or non-parametric, distribution-free as it is sometimes called. Bayesian methods necessarily involve a known distribution. We have described the concepts of Bayesian versus frequentist and parametric versus non-parametric in the introductory Chapters 1 and 2. Table 5.1 summarizes these apparent dichotomies and indicates appropriate usage. That non-parametric Bayesian tests do not exist appears self-evident, as the key Bayesian feature is the probability of a particular model in the face of the data. However, it is not quite this clear-cut, and there has been consideration of non-parametric methods in a Bayesian context (Gull & Fielden 1986). If we understand the data so that we can model its collection process, then the Bayesian route beckons (see Chapter 2 and its examples).
76
5.1 Methodology of classical hypothesis testing
77
Table 5.1. Usage of Bayes/frequentist/parametric/non-parametric testing Parametric
Non-parametric
Bayesian testing
Model known. Data gathering and uncertainty understood.
Such tests do not exist.
Classical testing
Model known. Underlying distribution of data known. Large enough numbers. Data on ordinal or interval scales.
Small numbers. Unknown model. Unknown underlying distributions or errors. Data on nominal or categorical scales.
And yet there are situations when classical methods are essential: • If we are comparing data with a model and we have very few of these data; or if we have poorly defined distributions or outliers then we do not have an adequate model for our data. We need non-parametric methods. • Classical methods are widely used. We therefore need to understand results quoted to us in these terms. The classical tests involve us in ‘rejecting the null hypothesis’, i.e. in rejecting rather than accepting a hypothesis at some level of significance. The hypothesis we reject may not be one in which we have the slightest interest. This is a process of elimination. A classical test works with probability distributions of a statistic while the Bayesian method deals with probability distributions of a hypothesis.
5.1 Methodology of classical hypothesis testing Classical hypothesis testing follows these steps. (1) Set up two possible and exclusive hypotheses, each with an associated terminal action: H0 , the null hypothesis or hypothesis of no effect, usually formulated to be rejected, and H1 , an alternative, or research hypothesis.
78
Hypothesis testing (2) Specify a priori the significance level α; choose a test which (a) approximates the conditions and (b) finds what is needed; obtain the sampling distribution and the region of rejection, whose area is a fraction α of the total area in the sampling distribution. (3) Run the test; reject H0 if the test yields a value of the statistic whose probability of occurrence under H0 is ≤ α. (4) Carry out the terminal action.
It is vital to emphasize (2). The significance level has to be chosen before the value of the test statistic is glimpsed; otherwise some arbitrary convolution of the data plus the psychology of the investigator is being tested. This is not a game; you must be prepared to carry out the terminal action on the stated terms. There is no such thing as an inconclusive hypothesis test! There are two types of error involved in the process, traditionally referred to (surprisingly enough) as Types I and II. A Type I error occurs when H0 is in fact true, and the probability of a Type I error is the probability of rejecting H0 when it is in fact true, i.e. α. The Type II error occurs when H0 is false, and the probability of a Type II error is the probability β of the failure to reject a false H0 ; β is not related to α in any direct or obvious way. The power of a test is the probability of rejecting a false H0 , or 1 − β. The sampling distribution is the probability distribution of the test statistic, i.e. the frequency distribution of area unity including all values of the test statistic under H0 . The probability of the occurrence of any value of the test statistic in the region of rejection is less than α, by definition; but where the region of rejection lies within the sampling distribution depends on H1 . If H1 indicates direction, then there is a single region of rejection and the test is one-tailed; if no direction is indicated, the region of rejection is comprised of the two ends of the distribution and we are dealing with a two-tailed test. This is the only use we make of H1 ; the testing procedure can only convince us to accept H1 if it is the sole alternative to H0 . The procedure of elimination serves to reject H0 , not prove H1 . Beware – it is human nature to think that your H1 is the only possible alternative to H0 . Both parametric and non-parametric (classical) tests follow this procedure; both use a test statistic with a known sampling distribution. The non-parametric aspect arises because the test statistic does not itself depend upon properties of the population(s) from which the data were
5.2 Parametric tests: means and variances, t and F tests
79
drawn. There are persuasive arguments for following non-parametric testing in using classical methods, as outlined at the head of Section 1.4. But first we consider the parametric route in some detail in order to establish methodology. 5.2 Parametric tests: means and variances, t and F tests A very common question arises when we have two sets of data (or one set of data and a model) and we ask if they differ in location or spread. The best-known parametric tests for such comparisons concern samples drawn from Normally distributed parent populations; these tests are of course the ‘Student’s’ t test (comparison of means) and the F test (comparison of variances), and are discussed in most books on statistics, e.g. Martin (1971), Stuart & Ord (1994). The t and F statistics have been introduced in Section 3.4. To contrast the classical and Bayesian methods for hypothesis testing, we look at the simple case of comparison of means. We deal with a Gaussian distribution, because its analytical tractability has resulted in many tests being developed for Gaussian data; and then, of course, there is the central limit theorem. Let us suppose we have n data Xi drawn from a Gaussian of mean µx , and m other data Yi , drawn from a Gaussian of identical variance but a different mean µy . Call the common variance σ 2 . The Bayesian method is to calculate the joint posterior distribution 2 2 1 i (xi − µx ) i (yi − µy ) prob(µx , µy , σ) ∝ n+m+1 exp − exp − σ 2σ 2 2σ 2 (5.1) in which we have used the Jeffreys prior (Exercise 2.6 of chapter 2) for the variance. Integrating over the ‘nuisance’ parameter σ, we would get the joint probability prob(µx , µy ) and could use it to derive, for example, the probability that µx is bigger than µy . From this we can calculate the probability distribution of (µx − µy ) (see e.g. Lee 1997, Chapter 5). The result depends on the data via a quantity (µx − µy ) − (X − Y ) √ t = (5.2) s m−1 + n−1 where nSx + mSy s2 = ν
80
Hypothesis testing with the usual mean squares Sx = (Xi − X)2 /n, similarly for Sy , and ν = n + m − 2. The distribution for t is −(ν+1)/2 Γ ν+1 t2 2 prob(t ) = √ . (5.3) 1+ ν πνΓ ν2 We regard the data as fixed and (µx − µy ) as the variable, simply computing the probability of any particular difference in the means. We might alternatively work out the range of differences which are, say, 90 per cent probable, or we might carry the distribution of (µx − µy ) on into a later probabilistic calculation. If we instead follow the classical line of reasoning, we do not treat the µ’s as random variables. Instead we guess that the difference in the averages X − Y will be the statistic we need; and we calculate its distribution on the null hypothesis that µx = µy . We find that X −Y t= √ s m−1 + n−1
(5.4)
follows a t distribution with n + m − 2 degrees of freedom. This is the classical Student’s t. Critical values are given in Table A2.3. This gives the basis of a classical hypothesis test, the t test for means. Assuming that (µx − µy ) = 0 (the null hypothesis), we calculate t. If it (or some greater value) is very unlikely, we think that the null hypothesis is ruled out. The t statistic is heavy with history and reflects an era when analytical calculations were essential. The penalty is the total reliance on the Gaussian. However, with cheap computing power we may expect to be able to follow the basic Bayesian approach outlined above for any distribution. By analogous calculations, we can arrive at the F test for variances. Again, Gaussian distributions are assumed. The null hypothesis is σx = σy , the data are Xi (i = 1, . . . , n) and Yi (i = 1, . . . , m) and the test statistic is (Xi − X)/(n − 1) F = i . i (Yi − Y )/(m − 1)
(5.5)
This follows an F distribution with n−1 and m−1 degrees of freedom (Table A2.4) and the testing procedure is the same as for Student’s t.
5.2 Parametric tests: means and variances, t and F tests
81
Clearly this statistic will be particularly sensitive to the Gaussian assumption.
Suppose we have two small sets of data, from Gaussian distributions of equal variance: −1.22, −1.17, 0.93, −0.58, −1.14 (mean −0.64), and 1.03, −1.59, −0.41, 0.71, 2.10 (mean 0.37), with a pooled standard deviation of 1.2. The standard t statistic is 1.12. If we do a two-tailed test (so being agnostic about whether one mean is larger than another), we find a 30 per cent chance that these data would arise if the means were the same. The one-tailed test (testing whether one mean is larger) gives 16 per cent. From a Bayesian point of view, we can calculate the distribution of (µx − µy ) for the same data. In Fig. 5.1 we can see clearly that one mean is smaller; the odds on this being so are about 10 to 1, as can be calculated by integrating the posterior distribution of the difference of means.
EXAMPLE
0.5 0.4 0.3 0.2 0.1 0 4
2
0 x
2
4
y
Fig. 5.1. The distribution of the difference of means for the example data.
5.2.1 The Behrens–Fisher test Relaxing the assumption of equal variances may be important. It is indeed possible to derive the distribution of the difference in means without the assumption of equal variances in the two samples; the resulting distribution is called the Behrens–Fisher distribution. It is of great interest in statistics because it is a rare example of a Bayesian analysis having no classical analogue; there is no classical test for the case of possibly unequal variances. Lee (1997) discusses this in some detail.
82
Hypothesis testing
The analytical form of the Behrens–Fisher distribution is complicated and involves a numerical integration anyway, so we may as well resort to a computer right away to calculate it from Bayes’ theorem. We suppose that our data are drawn from Gaussians with means µ and standard deviations σ. The joint posterior distribution (using the Jeffreys prior on the σ) is 2 1 i (xi − µx ) prob(µx , µy , σx , σy ) ∝ n+1 exp − 2σx2 σx (yi − µy )2 1 × n+1 exp − i . (5.6) 2σ 2 σy We have a multidimensional integration to do in order to get rid of the two nuisance parameters (σx and σy ) and to ensure that the resulting joint distribution prob(µx , µy ) is properly normalized. This is now not much of a problem, although until recently these integrations (for anything other than Gaussians) were a formidable obstacle to Bayesian methods. The analytical derivation of the Behrens–Fisher distribution eliminates all the numerical integrations bar one. Given the joint distribution of µx and µy , we would like the distribution of µy − µx . By changing variables we can easily see that ∞ prob(u = µy − µx ) = prob(v, v + u) dv. ∞
(Another integration!)
Consider the same example data as before, relaxing the assumption that the variances are equal. So although we cannot tell (classically) that the variances differ, we will obtain somewhat different results by not assuming that they are the same. We see from Fig. 5.2 that the distributions of µy − µx are very similar in either case, although as we might expect the distribution is a little wider if we do not assume that the variances are equal. The wings are broader and so tests are a little weaker (but may be more honest).
EXAMPLE
This general sort of Bayesian test can be followed for any distribution – as long as we know what it is, and can do the integrations.
5.2 Parametric tests: means and variances, t and F tests
83
0.5
prob
x
y
0.4 0.3 0.2 0.1 0 4
2
0 x
2
4
y
Fig. 5.2. Distribution of the difference of means assuming equal variances (dashed) and without this assumption (solid).
5.2.2 Non-Gaussian parametric testing In astronomy we frequently have little or no information about the distributions from which our data are drawn, yet we need to test whether they are the same or not. Since there is only one way in which two unknown distributions can be the same, but a multitude in which they may differ, it is not surprising that we currently have to work with classical hypothesis tests – ones which assume the distributions are the same. If we have some information about the distributions, we can use Bayesian methods. The trick here is to use a multiparameter generalization of a familiar distribution, where we carry the extra parameters to allow distortions in the shape. Eventually we can marginalize out these extra nuisance parameters, integrating over our prior assumptions about their magnitude. The most common example of this sort of generalization is the Gram– Charlier series:
x2 exp − 2 1+ ai Hi (x) (5.7) 2σ i in which the H’s are the Hermite polynomials. The coefficients ai are the free parameters we need. (Because the Hermite polynomials are orthogonal with respect to Gaussian weights, these coefficients are also related to the moments of the distribution we are trying to create.) The effect of these extra terms is to broaden and skew a Gaussian, and so for some data a few-term Gram–Charlier series may give quite a useful basis for a parametric analysis. Priors on the coefficients have to be set
84
Hypothesis testing 0.4
prob x
0.3 0.2 0.1 0 4
2
0 x
2
4
Fig. 5.3. Various distributions resulting from using just two terms in a Gram– Charlier distribution; the solid curve is a pure Gaussian.
by judgment. The even Hermite polynomials have the effect of changing scale, and so should follow the same Jeffreys prior as the standard deviation. The odd polynomials will change both scale and location and here setting the prior is less obvious. There are two other variants on the Gram–Charlier series. For a distribution allied to the exponential exp(−x/a), a Laguerre series will function in the same way as a Gram–Charlier series, except the distorting functions are the Laguerre polynomials. The Gamma series is based on the distribution xα (1 − x)β , defined on the interval from 0 to 1; the distorting functions are the even less familiar Jacobi polynomials. However, computer algebra packages such as MATHEMATICA give comprehensive support for special functions and make the application of these series rather straightforward (Reinking 2002). This approach clarifies the workings of non-parametric tests. Suppose we fix on a two-term Gram–Charlier expansion as a realistic representation of our data; the versatility is demonstrated in Fig. 5.3. For dataset (1) (1) 1, we then get the posterior prob(µ(1) , σ (1) , a1 , a2 ), and similarly for dataset 2. If we ask the apparently innocuous question ‘are these data drawn from different distributions?’ we see that there are many possibilities (in fact, 24 ) of the form, for instance, µ(1) > µ(2) and σ (1) < σ (2) (1) (2) (1) (2) and a1 > a1 and a2 < a2 . Working through these possibilities could be quite tedious. A different question might be ‘are these distributions at different locations, regardless of their widths?’, in which case we could marginalize out the σ’s and a2 ’s (Section 2.2); the location, in a Gram–Charlier expansion, is a simple combination of µ and a1 .
5.2 Parametric tests: means and variances, t and F tests
85
5.2.3 Which model is better? This does suggest that comparison of models in the sense ‘are these data drawn from the same distribution?’ might be a more tractable question. Notice that we are not asking if µ(1) = µ(2) , etc., as the probability of this event is zero. A useful way of answering this involves something called the Bayes factor or weight of evidence. Suppose we try to describe all of the data Xi , Yi with just one distribution G. This distribution may have parameters so let us denote this hypothesis by (G, θ). Alternatively (and by hypothesis exhaustively) we may use (Gx , θx ) for the data Xi and (Gy , θy ) for the data Yi . This hypothesis is (Gx , θx , Gy , θy ). Note we need prior probabilities for our two options, G or Gx Gy . Bayes’ theorem then tells us that prob(G, θ | X, Y ) =
prob(X, Y | G, θ)prob(G, θ) prob(G, θ | X, Y ) dθ +
prob(Gx , θx | X) dθ x
prob(Gy , θy | Y ) dθ y
(5.8)
in which the second term of the denominator arises because our alternative to (G, θ) is that the data are described as the product of two distinct distributions. The odds on the distinct distributions are (see Section 2.5)
prob(Gx , θx | X) dθx prob(Gy , θy | Y ) dθy
, prob(G, θ | X, Y ) dθ
(5.9)
and this ratio is closely related to the Bayes factor (see Lee 1997 for more details). To work out these odds we integrate the likelihood functions, weighted by the priors, over the range of parameters of the distributions.
Suppose we have the following two datasets: Xi = −0.16, 0.12, 0.44, 0.60, 0.70, 0.87, 0.88, 1.44, 1.74, 2.79 and Yi = 0.89, 0.99, 1.29, 1.73, 1.96, 2.35, 2.51, 2.79, 3.17, 3.76. The means differ by about one standard deviation. We consider two a-priori equally likely hypotheses. One is that all 20 data are drawn from the same Gaussian. The other is that they are drawn from different Gaussians. In the first case, the likelihood function is 2 (Yi − µ)2 1 i (Xi − µ) + √ exp − 2σ 2 ( 2πσ)20
EXAMPLE
86
Hypothesis testing
and we take the prior on σ to be σ1 . We also assume a uniform prior for the µ’s. In the second case, the likelihood is 2 2 1 1 i (Xi − µx ) i (Yi − µy ) √ √ exp − exp − 2σx2 2σy2 ( 2πσx )10 ( 2πσy )10 and the prior is σx1σy . Integrating over the range of the µ’s and σ’s, the odds on the data being drawn from different Gaussians are about 40 to 1 – a good bet. In the exercises we suggest following classical t and F tests on these data, and contrasting to the Bayes factor approach.
5.3 Non-parametric tests: single samples We now leave Bayesian methods and return to classical territory for the remainder of this chapter. ‘Non-parametric tests’ implies that ‘no distribution is assumed’. But let us not kid ourselves: something must be assumed, to make any progress. What is it? Various tests exploit different things, but a common method is to use counting probabilities. Take as an example the chi-square test (Section 5.3.1). The number of items in bin i is Ni , and we expect Ei . For smallish numbers, Poisson statistics tell us that the variance is also Ei . So (Ni − Ei )2 /Ei should be roughly a squared Gaussian variable, of unit variance. As another example, the runs test (Section 5.3.3) is just using the assumption that each successive observation is equally likely to be ‘up’ or ‘down’, so a Binomial distribution applies. The assumptions underlying non-parametric tests are weaker, and so more general, than for parametric tests. It is worth emphasizing again why we are going to advocate the nonparametric tests. • These make fewer assumptions about the data. If indeed the underlying distribution is unknown, there is no alternative. • If the sample size is small, probably we must use a non-parametric test. • The non-parametric tests can cope with data in non-numerical form, e.g. ranks, classifications. There may be no parametric equivalent. • Non-parametric tests can treat samples of observations from several different populations.
5.3 Non-parametric tests: single samples
87
What are the counter-arguments? The main one concerns binning – binning is bad; it loses information and therefore loses efficiency. The power of non-parametric tests may be somewhat less, but typically no more than 10 per cent less than their parametric equivalents. 5.3.1 Chi-square test Pearson’s (1900) paper in which chi-square was introduced is a foundation stone of modern statistical analysis1 ; a comprehensive and readable review (plus bibliography) is given by Cochran (1952). Consider observational data which can be binned, and a model/ hypothesis which predicts the population of each bin. The chi-square statistic describes the goodness-of-fit of the data to the model. If the observed numbers in each of k bins are Oi , and the expected values from the model are Ei , then this statistic is χ2 =
k
(Oi − Ei )2 i=1
Ei
.
(5.10)
The null hypothesis H0 is that the number of objects falling in each category is Ei ; the chi-square procedure tests whether the Oi are sufficiently close to Ei to be likely to have occurred under H0 . The sampling distribution under H0 of the statistic χ2 follows the chi-square distribution (Fig. 5.4) with ν = (k − 1) degrees of freedom. One degree of freedom is lost because of the constraint that Σi Oi = Σi Ei . The chi-square distribution is given by f (x) =
2−ν/2 ν/2−1 −x/2 e x Γ[ν/2]
(5.11)
(for x ≥ 0), the distribution function of the random variable Y 2 = Z12 + Z22 + . . . + Zν2 where the Zi are independent random variables of the standard Normal distribution. Table A2.6 presents critical values; if χ2 exceeds these values, H0 is rejected at that level of significance. 1
Pearson’s paper is entitled On the criterion that a given system of deviations from the probable in the case of a correlated system of variables is such that it can be reasonably supposed to have arisen from random sampling. It is wonderful polemic and gives several examples of the previous abuse of statistics, covering the frequency of buttercup petals to the incompetence of Astronomers Royal. (‘Perhaps the greatest defaulter in this respect is the late Sir George Biddell Airy.’) He demonstrates, for extra measure, that a run of bad luck at his roulette wheel, Monte Carlo, in July 1892 had one chance in 1029 of arising by chance; he avoids libel by phrasing his conclusion ‘. . . it will be more than ever evident how little chance had to do with the results . . . ’
88
Hypothesis testing
Fig. 5.4. The chi-square distribution: (a) f (χ2 , df), the probability density
∞ function of χ2 for df degrees of freedom; (b) the distribution function χ2 f (χ2 , df) d χ2 of Table A2.6, consulted to determine if χ2 is ‘large enough’ to reject H0 .
The premise of the chi-square test then is that the deviations from Ei are due to statistical fluctuations from limited numbers of observations per bin, i.e. ‘noise’ or Poisson statistics, and the chi-square distribution simply gives the probability that the chance deviations from Ei are as large as the observations Oi imply. As we shall see, we need enough data per bin to ensure that each term in the chi-square summation is approximately Gaussian. There is good news and bad news about the chi-square test. First the good: it is a test of which most scientists have heard, with which many are comfortable, and from which some are even prepared to accept the results. Moreover because χ2 is additive, the results of different datasets which may fall in different bins, bin sizes, or which may apply to different aspects of the same model, may be tested all at once. The contribution to χ2 of each bin may be examined and regions of exceptionally good or
5.3 Non-parametric tests: single samples
89
bad fit delineated. In addition, χ2 is easily computed, and its significance readily estimated as follows. The mean of the chi-square distribution equals the number of degrees of freedom, while the variance equals twice the number of degrees of freedom; see the plots of the function in Fig. 5.4. So as another rule of thumb, if χ2 should come out (for more than four bins) as ∼ (number of bins −1) then accept H0 . But if χ2 exceeds twice (number of bins − 1), probably H0 will be rejected. Finally minimizing χ2 is an exceptionally common method of model fitting (see Section 6.4); and an example of the chi-square test (and model fitting) is shown as Fig. 6.6. Now the bad news: the data must be binned to apply the test, and the bin populations must reach a certain size because it is obvious that instability results as Ei → 0. As another rule of thumb then: > 80 per cent of the bins must have Ei > 5. Bins may have to be combined to ensure this, an operation which is perfectly permissible for the test. However, the binning of data in general, and certainly the binning of bins, results in loss of efficiency and information, resolution in particular. Thus the advantages of the chi-square test are its general acceptance, the ease of computation, the ease of guessing significance, and the fact that model testing is for free. The disadvantages are the loss of power and information via binning, and the lack of applicability to small samples, in particular the serious instability at < 5 counts per bin. Moreover, the chisquare test cannot tell direction, i.e. it is a ‘ two-tailed’ test; it can only tell whether the differences between sample and prediction exceed those which can be reasonably expected on the basis of statistical fluctuations due to the finite sample size. There must be something better, and indeed there is: 5.3.2 Kolmogorov–Smirnov one-sample test The test is extremely simple to carry out: (i) Calculate Se (x), the predicted cumulative (integral) frequency distribution under H0 . (ii) Consider the sample of N observations, and compute So (x), the observed cumulative distribution, the sum of all observations to each x divided by the sum of all N observations. (iii) Find D = max|Se (x) − So (x)|
(5.12)
90
Hypothesis testing
(iv) Consult the known sampling distribution for D under H0 , as given in Table A2.7, to determine the fate of H0 . If D exceeds a critical value at the appropriate N , then H0 is rejected at that level of significance. Thus, as for the chi-square test, the sampling distribution indicates whether a divergence of the observed magnitude is ‘reasonable’ if the difference between observations and prediction is due solely to statistical fluctuations. The Kolmogorov–Smirnov test has some enormous advantages over the chi-square test. Firstly it treats the individual observations separately, and no information is lost because of grouping. Secondly, it works for small samples; for very small samples it is the only alternative. For intermediate sample sizes it is more powerful. Finally, note that as described here, the Kolmogorov–Smirnov test is non-directional or two-tailed, as is the chi-square test. However, a method of finding probabilities for the one-tailed test does exist (Birnbaum & Tingey 1951; Goodman 1954), giving the Kolmogorov–Smirnov test yet another advantage over the chi-square test. Then why not always use it? There are perhaps two valid reasons, in addition to the invalid one (that it is not so well known). Firstly the distributions must be continuous functions of the variable to apply the Kolmogorov–Smirnov test. The chi-square test is applicable to data which can be simply binned, grouped, categorized – there is no need for measurement on a numerical scale. Secondly, in model fitting and parameter estimation, the chi-square test is readily adapted (Section 6.4) by simply reducing the number of degrees of freedom according to the number of parameters adopted in the model. The Kolmogorov– Smirnov test cannot be adapted in this way, since the distribution of D is not known when parameters of the population are estimated from the sample.
5.3.3 One-sample runs test of randomness This delightfully simple test is contingent upon forming a binary (1–0) statistic from the sample data, e.g. heads-tails, or the sign of the residuals about the mean, or a best-fit line. It is to test H0 that the sample is random; that successive observations are independent. Are there too many or too few runs?
5.3 Non-parametric tests: single samples
91
Determine m, the number of heads or 1’s; n, the number of tails or 0’s, N = n + m; and r, the number of runs. Look up the level of significance from the tabled probabilities (Table A2.8) for a one- or two-tailed test, depending on H1 , which can specify (as the research hypothesis) how the non-randomness might occur. In general we are concerned simply with the one-tail test, asking whether or not the number of runs is too few, the issue being independence or otherwise of data in a sequence. Situations giving rise to too many runs are infrequent; but if indeed there are significantly too many runs it does say something serious about the data structure – probably in the sense that we do not understand it. In fact for m ‘heads’ and n ‘tails’ with N data, the expectation value of the number of runs is 2mn µr = (5.13) m+n+1 and in the large N approximation this is asymptotically Gaussian with 2nm(2nm − N ) σr = . (5.14) N 2 (N − 1) For large samples, then, it is possible to use the Normal distribution in the standard way by forming z=
r − µr σr
and consulting Table A2.1, the integral Gaussian or erf function. This is the procedure when the numbers exceed 20 and run off the end of Table A2.8.
Figure 5.5 shows the optical spectrum of quasar 3C207. The baseline has been estimated by the method of minimum Fourier components (Section 8.4.2). Does it fit properly? Is there low-level signal present in broad emission lines? Carefully selected regions of the spectrum are examined with the runs test. The runs test is applied by using one-bit digitization – is the datum above or below the fitted baseline? The lower-wavelength region has enhanced continuum, a quasar ‘blue bump’, where the likelihood of line emission is significantly reduced. The runs test yields concordance, 36 positive deflections, 29 negative, 31 runs against an expectation of
EXAMPLE
92
Hypothesis testing
Fig. 5.5. A spectrum of the quasar 3C207, taken with the 4.2-m William Herschel Telescope. The solid curve is a baseline fitted by a Fourier minimumcomponent technique. The regions considered for the runs test are shown in the separated sections, each with the baseline subtracted and magnified by a factor of 3.
32.1 runs, z = −0.28. The second region lies in the range of the hydrogen Balmer-line series, and several members are clearly present in emission. The result, a foregone conclusion here, is rejection of randomness by the runs test at about 4σ: 31 positives, 32 negatives, 16 runs against an expectation of 31.5, z = −3.94. The broad emission lines yield the contiguous regions that decrease the number of runs to a highly significant degree.
The test is at its most potent in looking for independence between adjacent sample members, e.g. in checking sequential data of scan or spectrum type as in the above example. It is frequently used for checking sequences of residuals, scatter of data about a model line, and in this guise it can give a straightforward answer as to whether a model is a good representation of the data.
5.4 Non-parametric tests: two independent samples Now suppose we have two samples; we want to know whether they could have been drawn from the same population, or from different
5.4 Non-parametric tests: two independent samples
93
populations, and if the latter, whether they differ in some predicted direction. Again assume we know nothing about probability distributions, so that we need non-parametric tests. There are several. 5.4.1 Fisher exact test The test is for two independent small samples for which discrete binary data are available, e.g. scores from the two samples fall in two mutually exclusive bins yielding a 2 × 2 contingency table as shown in Table 5.2. Table 5.2. 2 × 2 contingency table Sample = Category = 1 =2
1
2
A B
C D
H0 : the assignment of ‘scores’ is random. Compute the following statistic: (A + B)!(C + D)!(A + C)!(B + D)! . (5.15) N !A!B!C!D! This is the probability that the total of N scores could be as they are when the two samples are in fact identical. But in fact the test asks: What is the probability of occurrence of the observed outcome or one more extreme under H0 ? Hence by the laws of probability (see e.g. Stuart & Ord 1994), ptot = p1 + p2 + · · · ; computation can be tedious. Nevertheless this is the best test for small samples; and if N < 20, it is probably the only test to use. p=
5.4.2 Chi-square two-sample (or k-sample) test Again the much-loved chi-square test is applicable. All the previous shortcomings apply, but for data which are not on a numerical scale, there may be no alternative. To begin, each sample is binned in the same r bins (a k × r contingency table) – see Table 5.3. H0 is that the k samples are from the same population. Then compute χ2 =
r
k
(Oij − Eij )2 . Eij i=1 j=1
(5.16)
94
Hypothesis testing Table 5.3. Multi-sample contingency table Sample: j = Category: i = 1 2 3 4 5 .
1 O11 O21 O31 O41 O51 ...
2 O12 O22 O32 O42 O52 ...
3 O13 O23 O33 O43 O53 ...
The Eij are the expectation values, computed from k
Eij =
j=1
Oij .
r
Oij
i=1
r k
.
(5.17)
Oij
i=1 j=1
Under H0 this is distributed as χ2 , with (r−1)(k−1) degrees of freedom. Note that there is a modification of this test for the case of the 2 × 2 contingency table (Table 5.2) with a total of N objects. In this case, χ2 =
N (| AD − BC | −N/2)2 (A + B)(C + D)(A + C)(B + D)
(5.18)
has just one degree of freedom. The usual chi-square caveat applies – beware of the lethal count of 5, below which the cell populations should not fall in any number. If they do, combine adjacent cells, simulate the distribution of the test statistic under the null hypothesis or abandon the test. And if there are only 2 × 2 cells, the total (N ) must exceed 30; if not, use the Fisher exact probability test. There is one further distinctive feature about the chi-square test (and the 2 × 2 contingency-table test); it may be used to test a directional alternative to H0 , i.e. H1 can be that the two groups differ in some predicted sense. If the alternative to H0 is directional, then use Table A2.6 in the normal way and halve the probabilities at the heads of the columns, since the test is now one-tailed. For degrees of freedom > 1, the chi-square test is insensitive to order, and another test thus may be preferable.
5.4 Non-parametric tests: two independent samples
95
5.4.3 Wilcoxon–Mann–Whitney U test This test is usually preferable to χ2 , mostly because it avoids binning. There are two samples, A (m members) and B (n members); H0 is that A and B are from the same distribution or have the same parent population, while H1 may be one of three possibilities: (i) that A is stochastically larger than B; (ii) that B is stochastically larger than A; (iii) that A and B differ in some other way, perhaps in scatter or skewness. The first two hypotheses are directional, resulting in one-tailed tests; the third is not and correspondingly results in a two-tailed test. To proceed, first decide on H1 and of course the significance level α. Then (i) Rank in ascending order the combined sample A + B, preserving the A or B identity of each member. (ii) (Depending on the choice of H1 ) Sum the number of A-rankings to get UA , or vice versa, the B-rankings to get UB . Tied observations are assigned the average of the tied ranks. Note that if N = m+n, UA + UB =
N (N + 1) , 2
so that only one summation is necessary to determine both – but a decision on H1 should have been made a priori. (iii) The sampling distribution of U is known (of course, or there would not be a test). Table A2.9, columns labelled cu (upper-tail probabilities), presents the exact probability associated with the occurrence (under H0 ) of values of U greater than that observed. The table also presents exact probabilities associated with values of U less than those observed; entries correspond to the columns labelled cl (lower-tail probabilities). The table is arranged for m ≤ n, which presents no restriction in that group labels may be interchanged. What does present a restriction is that the table presents values only for m ≤ 4 and n ≤ 10. For samples up to m = 10 and n = 12, see Siegel & Castellan (1988). For still larger samples, the sampling distribution for UA tends to Normal with 2 mean µA = m(N +1)/2 and variance σA = mn(N +1)/12. Significance can be assessed from the Normal distribution, Table A2.1,
96
Hypothesis testing by calculating z=
UA ± 0.5 − µA σA
where +0.5 corresponds to considering probabilities of U ≤ that observed (lower tail), and −0.5 for U ≥ that observed (upper tail). If the two-tailed (‘the samples are distinguishable’) test is required, simply double the probabilities as determined from either Table A2.9 (small samples) or the Normal distribution approximation (large samples).
An application of the test is shown in Fig. 5.6, which presents magnitude distributions for flat and steep (radio) spectrum quasars from a complete sample of quasars in the Parkes 2.7-GHz survey (Masson & Wall 1977). H1 is that the flat-spectrum quasars extend to significantly lower (brighter) magnitudes than do the steepspectrum quasars, a claim made earlier by several observers. The eye agrees with H1 , and so does the result from the U test, in which we found U = 719, t = 2.69, rejecting H0 in favour of H1 at the 0.004 level of significance.
EXAMPLE
In addition to this versatility, the test has a further advantage of being applicable to small samples. In fact it is one of the most powerful non-parametric tests; the efficiency in comparison with the ‘Student’s’ t test is ≥ 95 per cent for even moderate-sized samples. It is therefore an obvious alternative to the chi-square test, particularly for small samples where the chi-square test is illegal, and when directional testing is desired. An alternative is the
5.4.4 Kolmogorov–Smirnov two-sample test The formulation parallels the Kolmogorov–Smirnov one-sample test; it considers the maximum deviation between the cumulative distributions of two samples with m and n members. H0 is (again) that the two samples are from the same population, and H1 can be that they differ (twotailed test), or that they differ in a specific direction (one-tailed test). To implement the test, refer to the procedure for the one-sample test (Section 5.3.2); merely exchange the cumulative distributions Se and So
5.4 Non-parametric tests: two independent samples
97
Fig. 5.6. Magnitude histograms for a complete sample of quasars from the Parkes 2.7-GHz survey, distinguished by radio spectrum. H0 , that the magnitude distributions are identical, is rejected using the Mann–Whitney–Wilcoxon U test at the 0.004 level of significance.
for Sm and Sn corresponding to the two samples. Critical values of D are given in Tables A2.10 and A2.11. Table A2.10 gives the values for small samples, one-tailed test, while Table A2.11 is for the two-tailed test. For large samples, two-tailed test, use Table A2.12. For large samples, one-tailed test, compute χ2 = 4D2
mn , m+n
(5.19)
which has a sampling distribution approximated by chi-square with two degrees of freedom. Then consult Table A2.6 to see if the observed D
98
Hypothesis testing
results in a value of χ2 large enough to reject H0 in favour of H1 at the desired level of significance. The test is extremely powerful with an efficiency (compared to the t test) of > 95 per cent for small samples, decreasing somewhat for larger samples. The efficiency always exceeds that of the chi-square test, and slightly exceeds that of the U test for very small samples. For larger samples, the converse is true, and the U test is to be preferred. Note that the Kolmogorov–Smirnov test can also be used to compare two-dimensional distributions (Peacock 1983).
Two examples, drawn from an investigation of flattening and radio emission among elliptical galaxies (Disney, Sparks & Wall 1984), are shown in Fig. 5.7. The upper diagrams compare the axial ratio b/a (minor to major axis) for (a) 102 bright ellipticals for which no radio emission was detected and (b) 30 ellipticals for which emission was detected. The Kolmogorov–Smirnov test rejects H0 , that the two distributions are from the same parent population, at the 1 per cent level of significance. The lower pair, to do with ascertaining whether seeing is affecting measurement of axial ratio (the radio ellipticals are on average more distant), shows some difference by eye, but no significant difference when the Kolmogorov–Smirnov test is carried out. These and tests on additional subsamples were used to show that there is a strong correlation between radio activity and flattening, in the sense that radio ellipticals are both inherently and apparently rounder than the average elliptical.
EXAMPLES
5.5 Summary, one- and two-sample non-parametric tests Tables 5.4, 5.5 and 5.6, adapted from Siegel & Castellan (1988), attempt a summary, demonstrating an apparent wide world of non-parametric tests available for sample comparison. But is this really so? In deciding which test(s), the following points should be noted; the decision may be made for you. (i) The two-sample and k-sample cases each contain columns of tests for related samples, i.e. matched-pair samples, or samples of paired replicates. This is common experimental practice in biological and
5.5 Summary, one- and two-sample non-parametric tests
99
Fig. 5.7. Kolmogorov–Smirnov tests on subsamples of ellipticals from the Disney–Wall (1977) sample of bright ellipticals. Upper panels – distribution functions in b/a, minor to major axis, for (a) the 102 undetected and (b) the 30 radio-detected ellipticals in the sample. The Kolmogorov–Smirnov twosample test rejects H0 , that the subsamples are drawn from the same population, at a significance level of < 1 per cent. Lower panels – distribution functions in log a/b for (c) the 51 ellipticals closer than 30 Mpc, (d) 76 bright ellipticals in the sample more distant than this. The Kolmogorov–Smirnov test indicates no significant difference between these latter subsamples.
behavioural sciences, where the concept of the control sample is highly developed. It is not so common in astronomy for obvious reasons, but has been exploited on occasion. The powerful tests available to treat such experiments are listed in Table 5.4, and are described by Siegel & Castellan. (ii) Table 5.4 runs downward in order of increasing sophistication of measurement level, from Nominal (in which the test objects are simply dumped into classes or bins) through Ordinal (by which objects are ranked or ordered) to Interval (for which objects are placed on a scale, not necessarily numerical, in which distance along the scale matters). None of the tests requires measurement on a Ratio scale, the strongest scale of measurement in which to
100
Hypothesis testing Table 5.4. Non-parametric tests for comparison of samples
Level of
One-sample
measurement case Nominal or categorical
Ordinal or ordered
Binomial test
Two–sample case Related McNemar change test
∗ Chi-square test
Independent ∗
Fisher exact test for 2 × 2 tables
k–sample case Related Cochran Q test
Independent ∗ Chi-square test for r × k tables
∗ Chi-square test for r × 2 tables
∗ Kolmogorov– Sign test Median test Smirnov one∗ U (Wilcoxon– sample test Wilcoxon signed-ranks Mann–Whitney) ∗ One-sample test test runs test Robust rankChange-point order test test ∗ Kolmogorov– Smirnov twosample test
Friedman two-way analysis of variance by ranks Page test for ordered alternatives
Extension of median test Kruskal– Wallis oneway analysis of variance Jonckheere test for ordered alternatives
Siegel–Tukey test for scale differences Interval
Permutation test for paired replicates
Permutation test for two independent samples Moses ranklike test for scale differences
∗
Described in this chapter; Siegel & Castellan (1988) discuss the other tests.
the properties of the interval scale a true zero point is added. (Degrees Celsius for temperature measurement represents an interval scale, and Kelvins a ratio scale.) An important feature of test selection lies in the level of measurement required by the test; the table is cumulative downward in the sense that at any level of measurement, all test above this level are applicable. (iii) The efficiency of a particular test depends very much on the individual application. Is the search for goodness-of-fit and general difference, i.e. is this sample from a given population? Are these
5.5 Summary, one- and two-sample non-parametric tests
101
Table 5.5. Single-sample non-parametric tests Test
Applicability†
N < 10? Comment
Binomial test
Goodness-of-fit (N )
Yes
Appropriate for two-category (dichotomous) data; do not dichotomize continuous data.
Goodness-of-fit (N )
No
For testing categorized, pre-binned, or classified data; choose categories with expected frequencies 6–10.
Kolmogorov– Goodness-of-fit Smirnov one- (O) sample test
Yes
The most powerful test for data from a continuous distribution; may always be more efficient than the chi-square test.
Yes
Does not estimate differences between groups.
∗
Chi-square test
∗
∗
One-sample runs test
Randomness of event sequences (O)
Change-point test
Change in the Yes distribution of an event sequence (O)
Robust with regard to changes in distributional form; efficient.
∗ †
Described in this chapter; Siegel & Castellan (1988) discuss the other tests. Goodness-of-fit indicates general testing for any type of difference, i.e. H0 is that the distribution is drawn from the specified population. The level of measurement required is indicated by N – Nominal, O – Ordinal, or I – Interval.
samples from the same population? Or is it a particular property of the distribution which is of interest, such as the location, e.g. central tendency, mean or median; or the dispersion, e.g. extremes, variance, rms. For instance in the two-sample case, the chi-square and the Kolmogorov–Smirnov (two-tailed) tests are both sensitive to any type of difference in the two distributions, location, dispersion, skewness, while the U test is reasonably sensitive to most properties, but is particularly powerful for location discrimination. To aid the process of choice, Tables 5.5 (single samples) and 5.6 (two samples) summarize the attributes of the one- and two-sample tests.
102
Hypothesis testing Table 5.6. Two-sample non-parametric tests
Test ∗
Fisher exact test for 2 × 2 tables
Applicability† N < 10? Comment Difference (N )
Yes
The most powerful test for dichotomous data.
Chi-square test for r × 2 tables
Difference (N )
No
Best for pre-binned, classified, or categorized data.
Median test
Location (O)
Yes
Best for small numbers; efficiency decreases with N .
U (Wilcoxon– Location Mann–Whitney) (O) test
Yes
One of the most efficient nonparametric tests.
Robust rankorder test
Yes
Efficiency similar to U test.
∗
∗
∗ Kolmogorov– Smirnov twosample test
Location (O)
Two-tailed: Yes Difference One-tailed: Location (O)
The most powerful test for data from a continuous distribution.
Siegel–Tukey test for scale differences
Dispersion (O)
Yes
The medians must be the same (or known) for both distributions. Low efficiency.
Permutation test
Location (I)
Yes
Very high efficiency.
(No)
Does not require identical medians; valid for small samples, but increases with sample size.
Moses rank-like Dispersion test for scale (I) differences ∗
Described in this chapter; Siegel & Castellan (1988) discuss the other tests. Difference signifies sensitivity to any form of difference between the two distributions, i.e. H0 is that the two distributions are drawn from the same population; Location indicates sensitivity to the position of the distributions, e.g. means or medians; and Dispersion indicates sensitivity to the spread of the distributions, i.e. variance, rms, extremes. The level of measurement required is indicated by N – Nominal, O – Ordinal, or I – Interval. †
Exercises
103
The choice of test may thus come down to Hobson’s. However, if it does not, and two (or more) alternatives remain, beware of this plot of the Devil. It might be possible to ‘test the tests’ in searching for support of a point of view. If such a procedure is followed, quantification of the amount by which significance is reduced must be considered: for a chosen significance level p in a total of N tests, the chance that one test will (randomly) come up significant is N p(1 − p)N −1 N p for small p. The application of efficient statistical procedure has power; but the application of common sense has more.
Exercises In the exercises denoted by (D), datasets are provided on the book’s website; or create your own. 5.1
5.2
5.3
5.4
Kolmogorov–Smirnov (D). Use the data provided, two datasets, one with a total of m = 290 observations, the other with 385 measurements. The former is of flux densities measured at random positions in the sky; the latter of flux densities at the positions of a specified set of galaxies. Using the Kolmogorov– Smirnov two-sample test, examine the hypothesis that there is excess flux density at the non-random positions. Wilcoxon–Mann–Whitney (D). Repeat the test with the Wilcoxon–Mann–Whitney statistic. Is the significance level different? How would you combine the results from these two tests, plus the chi-squared test in the text? t test and outliers (D). Create two datasets, one drawn from a Gaussian of unit variance, the other drawn from a variable combination of two Gaussians, the dominant one of unit variance and the other three times wider. All Gaussians are of zero mean. Perform a t test on sets of 10 observations and investigate what happens as contamination from the wide Gaussian is increased. Compare the effect on the posterior distribution of the difference of the means. Now shift the narrow Gaussian by half a unit, and repeat the experiment. What effect do the outliers have on our ability to refute the null hypothesis? How does the Bayesian approach compare? F test (D). Create some random data, as in the first part of Exercise 3. Investigate the sensitivity of the standard F test to a small level of contamination by outliers.
104 5.5
5.6
Hypothesis testing Non-parametric alternatives (D). Repeat the analysis of the last two exercises, using a non-parametric test; the Wilcoxon– Mann–Whitney test for the location test, and the Kolmogorov– Smirnov test for the variance test. How do the results compare with the parametric tests? Can you detect genuine differences in variance, apart from the outliers? Several datasets, one test. Suppose you have N independent datasets, and with a certain test you obtain a significance level of pi for each one. A useful overall significance is given by the W statistic (Peacock 1985) which is W =
N
pi .
i=1
5.7
5.8
Find the distribution of log W and describe how it could be used. Note this contrasts to the case discussed in the text, where we might perform several different tests on the same dataset. (Each pi will be uniformly distributed between zero and one, under the null hypothesis. The distribution of log W is the sum of these uniformly distributed numbers, and tends to a Gaussian of mean N and variance N .) Gram–Charlier (D). Take some data drawn from a Gaussian and investigate the posterior likelihood if just one term (the quadratic) is used in a Gram-Charlier expansion as an assumption for the ‘true’ distribution. Take the location as known. Find the distribution of the variance, marginalizing out the Gram– Charlier parameter. Also, find the odds on including the parameter in the model. What does this tell you about assuming a Gaussian distribution when the amounts of data are limited? Odds versus classical tests. Use the small dataset from the example in Section 5.2.3. Perform a classical analysis, using t and F tests. Compare and contrast to the odds calculated in the text. Does the Behrens–Fisher distribution give a better answer than either or both? See Jaynes’s comments on confidence intervals (Jaynes 1983).
6 Data modelling; parameter estimation
But what are the errors on your errors? (Graham Hine at a Mark Birkinshaw colloquium, Cambridge 1979)
Many pages of statistics textbooks are devoted to methods of estimating parameters, and calculating confidence intervals for them. For example, if our N data Zi follow a Gaussian distribution 2 1 z−µ prob(z) = √ , exp − 2σ 2 2πσ then the statistic m=
1 Zi N i
is a good estimator for µ and has a known distribution (a Gaussian again) which can be used for calculating confidence limits. Or, from the Bayesian point of view, we can calculate a probability distribution for µ, given the data. Any data-modelling procedure is just a more elaborate version of this, assuming we know the relevant probability distributions. Suppose our data Zi were measured at various values of some independent variable Xi , and we believed that they were ‘really’ scattered, with Gaussian errors, around the underlying functional relationship µ = µ(x, α1 , α2 , . . .), in which α1 , α2 , . . . are unknown parameters (slopes, intercepts, . . . ) of
105
106
Data modelling; parameter estimation
the relationship. We then have 1 (z − µ(x, α1 , α2 , . . .))2 prob(z | α1 , α2 . . .) = √ exp − 2σ 2 2πσ and, by Bayes’ theorem, we have the posterior probability distribution for the parameters 1 prob(α1 , α2 , . . . | Zi , µ) ∝ Πi √ 2πσ (Zi − µ(x, α1 , α2 , . . .))2 × exp − prob(α1 , α2 , . . .) 2σ 2
(6.1)
including as usual our prior information. We have included µ as one of the ‘givens’ to emphasize that everything depends on it being the correct model. This, at least formally, completes our task; we have a probability distribution for the parameters of our model, given the data. This is a very general approach. In the limiting case of uninformative or diffuse priors, it is very closely related to the method of maximum likelihood; if the distribution of the residuals from the model is indeed Gaussian, it is closely related to the method of least squares. Moreover, it can be used in a clear way to update models as new data arrive; the posterior from one stage of the experiments becomes the prior for the next. We can also deal nicely with unwanted parameters (‘nuisance’ parameters). Typically we will end up with a probability distribution for various parameters, some of interest (say, cosmological parameters) and some not (say, instrumental calibrations). We can marginalize out the unwanted parameters by an integration, leaving us with the distribution of the variable of interest that takes account of the range of plausible values of the unwanted variables. Later examples will develop these ideas. Modelling can be a very expensive part of any investigation. Analytic approximations were developed in past years for very good reasons. Modelling processes always involve finding an extreme value, a maximum or minimum, of some merit function. Without help from an analytic solution, this means evaluating the function, and perhaps its derivatives, many times. The model itself may be the result of a complex and timeconsuming computation, so evaluating it over a range of parameters is even worse. Another difficulty that arises in the Bayesian approach is numerical integration. Interesting problems have many parameters; marginalizing
6.1 The maximum-likelihood method
107
these out, or calculating evidences for discriminating between models, involves multidimensional integrals. These are often very time-consuming, and laborious to check. Any analytical help we can get is especially welcome in doing integrations. We will see the relevance of this in the next section, where powerful theorems may allow great simplifications. Perhaps the most important thing to remember about models is blindingly obvious; they may be wrong. The most insidious case of this is a mistake in the assumed distribution of residuals about the model. Inevitably, the parameters deduced from the model will be wrong. Worse, the inferred errors on these parameters will be wrong too, often giving a quite false sense of security. It is important to have a range of models available, and always to check optimized models against the data, inspecting the residuals for strange outliers or clusters of positive or negative residuals. The runs test (Section 5.3.3) is helpful in this respect.
6.1 The maximum-likelihood method Maximum likelihood (ML) has a long history: it was derived by Bernoulli in 1776 and Gauss around 1821, and worked out in detail by Fisher in 1912. We have met the likelihood function several times already; together with the prior probabilities, it makes up the posterior probability from Bayes’ theorem. Suppose our data are described by the probability density function f (X; α), where x is a variable, and α is a parameter (maybe many parameters) characterizing the known form of f . We want to estimate α. If X1 , X2 , . . . , XN are data, presumed independent and all drawn from f , then the likelihood function is L(X1 , X2 , . . . , XN ) = f (X1 , X2 , . . . , XN | α) = f (X1 | α)f (X2 | α) . . . f (XN | α) =
N
f (Xi | α).
(6.2)
From the classical point of view this is the probability, given α, of obtaining the data. From the Bayesian point of view it is proportional to the probability of α, given the data and assuming that the priors are ‘diffuse’. Practically speaking, this means that they change little over the peaked region of the likelihood function. Finding the constant of proportionality involves the troublesome integrals we referred to before.
108
Data modelling; parameter estimation
If the priors are not diffuse, this means they are having as strong an effect on our conclusions as the data. This is not an unlikely situation, but it does rule out the handy analytical approximations we will describe later. From either point of view, more intelligibly from the Bayesian, the peak value of L seems likely to be a useful choice of the ‘best’ estimate of α. This does rather depend on what we want to do next with our estimate, however. Formally, the maximum-likelihood estimator (MLE) of α is α ˆ = (that value of α which maximizes L(α) for all variations of α). Often we can find this from ∂ ln L(α) =0 (6.3) ∂α α=α ˆ but sometimes we cannot – an example of this will be given later. Maximizing the logarithm is often convenient, both algebraically and numerically. The MLE is a statistic – it depends only on the data, and not on any parameters.
Consider our old friend the regression line, for which we have values of Yi measured at given values of the independent variable Xi . Our model is
EXAMPLE
y(a, b) = ax + b and assuming that the Yi have a Gaussian scatter, each term in the likelihood product is (Yi − (aXi + b))2 Li (y|(a, b)) = exp − 2σ 2 i.e. the residuals are (Yi −model), and our model has the free parameters (a, b). Maximising the log of the likelihood products then yields ∂Σ = −2Σ(Yi − a − bXi ) = 0 ∂a ∂Σ = −2ΣXi (Yi − a − bXi ) = 0 ∂b from which two equations in two unknowns we get the well-known a=
ΣYi (Xi − X) 2
Σ(Xi − X )
b = Y − aX.
6.1 The maximum-likelihood method
109
With this simple maximum-likelihood example, we have accidentally derived the standard OLS, the ordinary least squares estimate of y on the independent variable x. But note how this happened: we were given the fact that the Yi were Normally distributed with their scatter described by a single deviation σ; and of course we were given the fact that a straight-line model was correct. It need not be this way: we could have started knowing that each Yi had an associated σi , or even that the distribution in y about the line was not Gaussian, perhaps say uniform, or dependent on |Yi − model| rather than (Yi − model)2 . The formulation is identical, although the algebra may not work out as neatly as it does for an OLS regression line. But this of course is another advantage of maximum likelihood – the likelihood function can be computed and the maximum found without recourse to algebra.
Jauncey (1967) showed that maximum likelihood was an excellent way of estimating the slope of the number–flux-density relation, the dependence of source surface density on intensity, for extragalactic radio sources. The source count is assumed to be of the power-law form
EXAMPLE
N (> S) = kS −γ where N is the number of sources on a particular patch of sky with flux densities greater than S, k is a constant, and −γ is the exponent, or slope in the log N – log S plane, which we wish to estimate; see Fig. 6.1. The probability distribution for S (the chance of getting a source with a flux density near S) is then prob(S) = γkS −(γ+1) and k is determined by the normalization to unity
∞ prob(S) dS = 1. S0
(We have taken the maximum possible flux density to be infinity, with small error for steep counts.) k is then γ/S0γ and the log-likelihood is, dropping constants, Si ln L(γ) = M ln γ − γ ln S0 i where we have observed M sources with flux densities S brighter than
110
Data modelling; parameter estimation
Fig. 6.1. A maximum-likelihood application. The figures show differential source counts generated via Monte Carlo sampling with an initial uniform deviate (see Section 6.5) obeying the source-count law N (> S) = kS −1.5 . The straight line in each shows the anticipated count with slope −2.5: left, k = 1.0, 400 trials; right, k = 10.0, 4000 trials. The ML results for the slopes are −2.52 ± 0.09 and −2.49 ± 0.03, the range being given by the points at which the log likelihood function has dropped from its maximum √ by a factor of 2. The anticipated errors in the two exponents, given by |slope|/ trials (see the next-but-one example), are 0.075 and 0.024.
S0 . Differentiating this with respect to γ to find the maximum then gives the equation for γˆ , the MLE of γ: M Si i ln S0
γˆ =
a nicely intuitive result. This application of ML makes optimum use of the data in that the sources are not grouped and the loss of power which always results from binning is avoided.
The MLE cannot always be obtained by differentiation, as the following example shows.
6.1 The maximum-likelihood method
111
A supernova produces an intense burst of neutrinos. The intensity of this burst decays exponentially after the core collapse of the precursor star. A handful of neutrinos (say N in number) were detected from supernova 1987, with arrival times (in order) T1 , T2 , . . . The probability of a neutrino arriving at time t is
EXAMPLE
prob(t) = exp[−(t − t0 )] for t > t0 and zero otherwise. Times are measured in units of the half-life and t0 is the parameter we want, the start of the burst. The log-likelihood is just ln L(t0 ) = N t0 − Ti i
and this doesn’t appear to have a maximum. However, clearly t0 < T1 and so the likelihood is maximized, within the allowable range of t0 , at tˆ0 = T1 .
After the MLE estimate has been obtained, it is essential to perform a final check: does the MLE model fit the data reasonably? If it does not then the data are erroneous when the model is known to be right; or, the adopted or assumed model is wrong; or (most commonly) there has been a blunder of some kind. There are many ways of carrying out such a check; two of these, the chi-square test and the Kolmogorov–Smirnov test, were described in Sections 5.3.1 and 5.3.2, respectively. If the deviations between the best-fit model and the data (the residuals) are Gaussian, the log-likelihood function becomes a sum of squares of residuals and we have the famous method of least squares. More on this later. Now for those theorems. The strongest reason for picking the MLE of a parameter is that it has desirable properties – it has minimum variance compared to any other estimate, and it is asymptotically Normally distributed around the true value. An MLE is not always unbiased, however. If we estimate a vector α ˆ by the maximum-likelihood method, then the components of the estimated vector are asymptotically distributed around the true value like a multivariate Gaussian (Section 4.2). ‘Asymptotically’ means when we have lots of data, strictly speaking infinite
112
Data modelling; parameter estimation
amounts. The covariance matrix that describes this Gaussian can be derived from the second derivatives of the likelihood with respect to the parameters. This involves a famous matrix called the Hessian, which is ∂ 2 ln L ∂ 2 ln L ∂ 2 ln L ∂α1 ∂α2 ∂α1 ∂α3 . . . ∂α21 ∂ ln L ∂ 2 ln L ∂ 2 ln L ∂α ∂α 22 1 2∂α22 ∂α22 ∂α3 . . . H = ∂ ln L ∂ ln L ∂ ln L (6.4) . ∂α3 α1 ∂α3 ∂α2 ∂α2 . . . 3 .. .. .. . . . This matrix of course depends on the data. Taking its expectation value (the ‘average’ value of each component of the matrix, E[H] for short, Section 3.1), we have a simple expression for the covariance matrix of the multivariate Gaussian distribution of the maximum-likelihood estimators of the parameters: C = (E[H])−1 ,
(6.5)
the (. . .)−1 signifying the inverse matrix. The probability distribution of our N MLEs α ˆ is then prob(αˆ1 , αˆ2 , . . .) =
1 (2π)N
1 ˆ − α) · C −1 · (α exp − (α ˆ − α)T 2 |C| (6.6)
ˆ is distributed around the true value α so that, as stated, the MLE (α) with a spread described by the covariance C. | C | is the determinant of C. Taking the expectation value is obviously important, as otherwise the matrix would be different for each set of data. Sometimes we can carry out the expectation, or averaging, operation analytically in terms of α , the parameters of the original model. Sometimes the matrix does not involve the data at all. Most commonly, we just have to take the single matrix, given by our one set of data, as the best estimate we can make of the average value. Why should the maximum-likelihood estimators obey this theorem? Take a simple case, a Gaussian of true mean µ and variance σ 2 . If we
6.2 The method of least squares: regression analysis
113
have N data Xi , the log likelihood is (dropping constants) −1 log L = (Xi − µ)2 − N log σ 2σ 2 i and −∂ 2 log L N = 2. ∂µ2 σ This is the Hessian ‘matrix’ for our simple problem. Taking its expectation and then inverse, not too hard in this case, gives us the variance on the estimate of the mean as σ 2 /N , the anticipated result. This example provides some justification for the theorem. In the exercises we set the somewhat more complicated case of estimating µ and σ together. This gives a matrix problem rather than a scalar one, and some real expectations have to be performed.
In the source-count example, we have just one parameter. The variance on γˆ is then −1 ∂ 2 L(γ) E ∂γ 2
EXAMPLE
which is γ 2 /M , the expectation is easy in this case. However, we see that the error is given in terms of the thing we want to know, namely γ. As long as the errors are small we can approximate them by γˆ 2 /M.
6.2 The method of least squares: regression analysis Least squares is a famous old method of dealing with noisy data; it was invented, for astronomical use, by Gauss and Laplace at the beginning of the nineteenth century. There is a huge literature, e.g. Williams (1959); Linnik (1961); Montgomery & Peck (1992). The justification for the method follows immediately from the method of maximum likelihood; if the distribution of the residuals is Gaussian, then the log likelihood is a sum of squares of the form log L = constant −
N i=1
ξi (Xi − µ(α1 , α2 , . . .))2
(6.7)
114
Data modelling; parameter estimation
where the ξ are the weights, obviously inversely proportional to the variance on the measurements. Usually the weights are assumed equal for all the data, and least squares is just that; we seek the model parameters which minimize log L = constant −
N 1 (Xi − µ(α1 , α2 , . . .))2 . 2σ 2 i=1
These will just be the maximum-likelihood estimators, and everything we have said before about them carries over. In particular, they are asymptotically distributed like a multivariate Gaussian. If we do not know the error level (the σ) we do not need to use it, but we will not be able to infer errors on the MLE; we will get a model fit, but we will never know how good or bad the model is. The matrix of second derivatives defining the covariance matrix of the estimates, the Hessian matrix (Section 6.1), takes on a particular significance in the method of least squares because it is often used by the numerical algorithms which find the minimum. There are many powerful variations on these algorithms – see Numerical Recipes (Press et al. 1992) for details. Typically the value of the Hessian matrix, at the minimum, pops out as a by-product of the minimization. We can use this directly to work out the covariance matrix, as long as our model is linear in the parameters; in this case, the expectation operation is straightforward and the matrix does not depend on any of the parameters. We saw before why this is a problem (in the source-count example) – we want to find the parameters, and using the estimates in the covariance matrix is not an ideal procedure. The notion of a linear model is worth clarifying. Suppose our data Xi are measured as a function of some independent variable Zi . Then a linear model – linear in the parameters – might be αz 2 + β exp(−z), whereas α exp(−βz) is not a linear model. Of course a model may be approximately linear near the MLE. However, how close must it be? This illustrates again the general feature of the asymptotic Normality of the MLE – we can use the approximation, but we can’t tell how good it is. Usually things will start to go wrong first in the wings of the inferred distributions (we have seen this in a previous example) and so high degrees of significance usually cannot be trusted unless they have been calculated exactly, or simulated by Monte Carlo methods.
6.2 The method of least squares: regression analysis EXAMPLE
115
In the notation we used before, suppose our model is µ(α, β) = αz + βz 2 ,
a simple polynomial. The covariance matrix can be calculated from H, the matrix of derivatives of the log likelihood; it is just 2 1 Zi 0 i C= 2 4 0 σ i Zi so the variance on β, for example, is σ 2 / i Zi4 . Evidently where we make the measurements (the Zi ) will affect the variance. The effects are obvious enough in this simple case, but in more complicated cases it may be worth examining the experimental design, via the covariance matrix, to minimize the expected errors.
Quite often we will not be confident that we are dealing with Gaussian residuals, and usually this is because of outliers – residuals which are extremely unlikely on the Gaussian hypothesis. One convenient distribution which has ‘fat’ tails, and is a useful contrast to a Gaussian, is the simple exponential 1 |x−µ| prob(x) = exp − . 2a a If the residuals are distributed in this way, then it is easy to see that maximum likelihood leads to the minimization of the sum of the absolute values of the residuals. A t distribution may also be a helpful model. Working out a MLE in this way will give some indication of whether outliers are driving the answer. The only problem may be that relatively slow numerical routines have to be used; least squares minimization routines are highly developed by comparison. Let us return for the last time to our simple regression line, the least squares fit of the model y = ax + b through N pairs of (Xi , Yi ) by minimizing the squares of the residuals. This yields the well-known expressions for slope and intercept (differing slightly from those in the first example of Section 6.1, but readily shown to be equivalent): a=
N
N
N
Xi Yi −
N
Xi2
−
N
Xi
N
N
Xi
Yi 2
(6.8)
116
Data modelling; parameter estimation
and b=
N
Yi − a
N
Xi
/N.
(6.9)
In the absence of knowledge of the how and why of a relation between the Xi and the Yi (Section 4.4), any two-parameter curve may be fitted to the data pairs just with simple coordinate transformations; for example (i) an exponential, y = b exp a requires Yi to be changed to ln Yi in the above expressions, (ii) a power-law, y = bxa ; change Yi to ln Yi and Xi to ln Xi ; √ (iii) a parabola, y = b + ax2 ; change Xi to Xi . Note that the residuals cannot be Gaussian for all of these transformations (and may not be Gaussian for any): of course it is always possible to minimize the squares of the residuals, but it may well not be possible to retain the formal justification for doing so. The tests of Chapter 4 can be revealing as to which (if any) model fits, particularly the runs test. This simple formulation of the least squares fit for y on x represents the tip of an iceberg – there is an enormous variety of least squares linear regression procedures. Amongst the issues involved in choosing a procedure: • Are the data to be treated weighted or unweighted? • (And the related question) Do all the data have the same properties, e.g. in the simple case of y on x, is one σy2 applicable to all y? Or does σy2 depend on y? In the uniform σ case, the data are described as homoskedastic, and in the opposite case, heteroskedastic. • Is the right fit the standard ordinary least squares solution y on x (OLS(Y /X)) or x on y (OLS(X/Y )? Or something different, as discussed below? • If we know we have heteroskedasticity, with the uncertainty different but known in each yi and perhaps also in each xi , how do we use this information to estimate the uncertainty in the fit? • Are the data truncated or censored; do we wish to include upper limits in our fit? This is perfectly possible; see Section 7.5. The thorough papers of Feigelson and collaborators (Isobe et al. 1990; Babu & Feigelson 1992; Feigelson & Babu 1992b) consider these issues, describe the complexities, indicate how to find errors with bootstrap
6.2 The method of least squares: regression analysis
117
and jackknife resampling (Section 6.6), and identify appropriate software routines. In the astronomical context, Feigelson & Babu (1992b) emphasize that much of the proliferation of linear regression methods in the cosmic distance-scale literature is due to lack of precision in defining the scientific question. The question defines the statistical model. The serious fitter must consult the Feigelson references. In the interim and as an indication of why you must, consider the following example.
Return to our bivariate Gaussian of Section 4.2 and Fig. 4.4, and now consider random variates (xi , yi ) selected (a) in accord with ρ = 0.05 (little correlation) and ρ = 0.95 (strongly correlated). The ellipses of the contours are shown in Fig. 6.2. For the case of little correlation, the two OLS lines are stunningly different, almost orthogonal; for the relatively strong correlation, the lines are very similar.
EXAMPLE
Fig. 6.2. Linear contours of the bivariate Gaussian probability distribution. Left: ρ = 0.05, a bivariate distribution with weak connection between x and y; right: ρ = 0.95, indicative of a strong connection between x and y. In each case 5000 (x, y) pairs have been plotted, selected at random from the appropriate distribution as described in Section 6.5. Two lines are shown as fits for each distribution, the OLS(X/Y ) and the OLS(Y /X).
The point is that we know the answer here for the relation: it is a line of slope unity, 45◦ . With little (yet formally significant) correlation, the OLS lines mislead us dramatically. Of course the so-called bisector
118
Data modelling; parameter estimation
line (the average of the two OLSs) would get it right, as would the orthogonal regression line which minimizes the perpendicular distances. But for the former, if the points were not Gaussian in distribution, would you trust it? A few outliers (mistakes?) would soon wreck it. The latter is principal component analysis (Section 4.5) precisely. It has already been emphasized that when the dependences of variables on each other are not understood, PCA is the way to go. It gives the right answer in this example; it tells us what the relation between y and x is, without us assuming which variable is in control. It is the right answer if we want to describe a relation between x and y. So far, we have followed classical lines in our discussion of likelihood. The method is attractive and very useful; the main limitation is the difficulty in calculating the parameters of the asymptotic distribution of the MLE. And, of course, without an exact solution it is difficult to be sure how useful this asymptotic distribution is anyway.
6.3 Bayesian likelihood analysis Bayes’ theorem says, for model parameters (a vector, in general) α and data Xi , prob( α | Xi ) ∝ L( α | Xi )prob( α)
(6.10)
so the likelihood function is important here too. However, given the posterior probability of α , we may choose to emphasize properties other than the most probable α – we may only be interested in the probability that it exceeds a certain value, for example. Two great strengths of the Bayesian approach are the ability to deal with nuisance parameters via marginalization, and the use of the evidence or Bayes factor to choose between models. Another useful product of the Bayesian approach is the asymptotic distribution of the likelihood function itself. L( α) is asymptotically a multivariate Gaussian distributed around the MLE α ˆ , with covariance matrix given by the inverse of ∂ 2 ln L ∂ 2 ln L ∂ 2 ln L ∂α1 ∂α2 ∂α1 ∂α3 . . . ∂α21 ∂ ln L ∂ 2 ln L ∂ 2 ln L ∂α ∂α 22 1 2∂α22 ∂α22 ∂α3 . . . − ∂ ln L ∂ ln L ∂ ln L (6.11) . ∂α3 α1 ∂α3 ∂α2 ∂α2 . . . 3 .. .. .. . . . evaluated at the peak, namely the MLE of α .
6.3 Bayesian likelihood analysis
119
We will illustrate this approach by developing a simple two-parameter example, fitting a power law to some radio flux-density data. This example will appear in various guises in this chapter, but each time we will assume Gaussian statistics and uniform, or diffuse priors. These assumptions do not simplify the calculations, which were all done numerically in any case; they do simplify the presentation. Use the error distribution and prior that fits your problem.
Let us suppose we have flux density measurements at 0.4, 1.4, 2.7, 5 and 10 GHz. The corresponding data are 1.855, 0.640, 0.444, 0.22 and 0.102 flux units – see Fig. 6.3. Let us label the frequencies as fi and the data as Si . These follow a power law of slope −1, but have a 10 per cent Gaussian noise added. The noise level is denoted , and the model for the flux density as a function of frequency is kf −γ . Assuming we know the noise level and distribution, each term in the likelihood product is of the form 2 − Si − kfi−γ 1 √ . exp 2 2π kfi−γ 2 kfi−γ EXAMPLE
0.5 0.25 0 − 0.25 − 0.5 − 0.75 −1 − 1.25 − 0.4− 0.2 0 0.2 0.4 0.6 0.8 1 Log frequency
Log flux
Log flux
The likelihood is therefore a function of k and γ. A contour map of the log likelihood is in Fig. 6.4. We can calculate the Gaussian approximation to the likelihood, also shown in Fig. 6.4. At this point, there are at least two possibilities for further analysis. We may wish to know which pairs of (k, γ) are, say, 90 per cent probable. This in general involves a very awkward integration of the posterior probabilities. The multivariate Gaussian approximation to the likelihood is much easier to use; it is automatically normalized and there are analytic forms for its integral over any number of its arguments (see, for example, Jaynes 2003). As 0.5 0.25 0 − 0.25 − 0.5 − 0.75 −1 − 1.25 − 0.4− 0.2 0 0.2 0.4 0.6 0.8 1 Log frequency
Fig. 6.3. The two experimental spectra we will examine; the right-hand one contains an offset error as well as random noise.
120
Data modelling; parameter estimation
1.4
prob K
K
1.2 1 0.8
7 6 5 4 3 2 1 0
0.6 0.4
0.6
0.8
1
1.2
0.7
0.75
0.8
0.85 K
0.9
0.95
1
0.7
0.75
0.8
0.85
0.9
0.95
1
1.4
1.4 8 1.2 prob
K
6 1
4 2
0.8
0 0.6 0.4
0.6
0.8
1
1.2
1.4
Fig. 6.4. Top left, a contour plot of the log likelihood function; bottom left, the Gaussian approximation; right panels, the marginal distributions of k and γ, comparing the Gaussian approximation to the full likelihood.
can be seen in the figure, the areas defined by a particular probability requirement are simple ellipses. Another possibility is to ask for the probability of, say, k regardless of γ. So we have a posterior probability prob(k, γ | Si ) and we form
prob(k | Si ) = prob(k, γ | Si ) dγ. The probability distributions for k and γ are also shown in Fig. 6.4, along with the distributions deduced from the Gaussian approximation. As we can see the agreement is quite good.
6.3 Bayesian likelihood analysis
121
Marginalization (Section 2.2) can be a very useful technique. Often we are not interested in all the parameters we need to estimate to make a model. If we were investigating radio spectra, for instance, we would want to marginalize out k in our example. We may also have to estimate instrumental parameters as part of our modelling process, but at the end we marginalize them out in order to get answers which do not depend on these parameters. Of course, the marginalization process will always broaden the distribution of the parameters we do want, because it is absorbing the uncertainty in the parameters we don’t want – the nuisance parameters.
In our radio spectrum example (Fig. 6.3) we will add (somewhat artificially) an offset of 0.4 flux units to each measurement. This has the effect of flattening the spectrum quite markedly. We will calculate two possibilities. Model A is the simple one we assumed before, with no offsets built in. Model B uses a model for the flux densities of the form β + kf −γ . Each likelihood term is then
EXAMPLE
2 − Si − β + kfi−γ 1 √ . exp 2π kfi−γ 2( kfi−γ )2 We also suppose that we have some suspicion of the existence of this offset, so we place a prior on β of mean 0.4, standard deviation . Model B therefore returns a posterior distribution for k, γ and β. We are not actually interested in β (although an instrumental scientist might be) so we marginalize it out. The likelihoods from the two models are shown in Fig. 6.5, and it is clear that the more complex model does a better job of recovering the true parameters. The procedure works because there is information in the data about both the instrumental and the source parameters, given the model of the spectrum. If our model for the spectrum had a ‘break’ in it, we would not be able to recover much information about β, if any. If our fluxes had a pure scale error, we would not have been able to recover this either.
In the real world, of course, we do not have the truth available to guide us as to our choice of model A or model B. As remarked before, we ought to check the ‘fit’ of the two models. In one dimension there are
122
Data modelling; parameter estimation
1.4
K
1.2
1
0.8
0.6 0.4
0.6
0.8
1
1.2
1.4
Fig. 6.5. The log likelihoods for the two models; the black contours are for model A and the dashed contours are for model B.
various ways to do this, as discussed in Chapter 4. In many dimensions things are harder. At the risk of repetition, let’s look again at the use of evidence (the Bayes factor). Suppose we are choosing between model A and model B and we believe they are the only possibilities. The prior probability of A is, say, pA and of B is pB . The posterior probability of the parameters α, given data Xi , is prob(α | Xi , A, B) pA L(Xi | α, A)prob(α | A) + pB L(Xi | α, B)prob(α | B) = prob(Xi )
(6.12)
where we are emphasizing which model enters the various likelihoods. prob(Xi ) is the normalizing factor which ensures that the posterior distribution is properly normalized; its calculation usually involves a multidimensional integral. prob(α | A) is the prior on α in model A, and similarly for B. The posterior odds on model A, compared to model B, are then simply p L(Xi | α, A)prob(α | A) α A (6.13) p L(Xi | α, B)prob(α | B) α B in which we have to integrate over the range of parameters appropriate to each model. This is worth the effort because we get a straightforward answer to the question: which of A or B would it be better to bet on?
6.4 The minimum chi-square method
123
In the previous two examples we have worked out the likelihood functions, which we abbreviate L(Xi | k, γ, A) for model A and similarly for model B. In model B we also have a prior on the offset β, which is 1 −(β − 0.4)2 prob(β | B) = √ . exp 2( )2 2π
EXAMPLE
We then form the ratio of the integrals
pA dk dγL(Xi | k, γ, A) and
pB
dk
dγ
dβL(Xi | k, γ, B)prob(β | B).
Let’s take pA = pB , an agnostic prior state; note we have implicitly assumed uniform priors on k and γ. Cranking through the integrations numerically, we get: odds on B compared to A: about 8 to 1. Another way of looking at this is that we would have had to have been prepared to offer prior odds of 8:1 against the existence of the offset, for the posterior odds to have been even.
6.4 The minimum chi-square method Yet of course there are occasions when Bayesian methods fail us – perhaps we have been given the data in binned form, or indeed somebody else has used classical modelling methods which we wish to examine. A dominant classical modelling process is minimum chi-square, a simple extension of the chi-square goodness-of-fit test described in Section 4.2. It will be seen that it is closely related to least squares and weighted least squares methods, and in fact the minimum chi-square statistic has asymptotic properties similar to ML. Consider observational data which can be (or are already) binned, and a model and hypothesis which predicts the population of each bin. The chi-square statistic describes the goodness-of-fit of the data to the model. If the observed numbers in each of k bins are Oi , and the expected
124
Data modelling; parameter estimation
values from the model are Ei , then this statistic is 2
χ =
k (Oi − Ei )2 i=1
Ei
.
(6.14)
(The parallel with weighted least squares is evident: the statistic is the squares of the residuals weighted by what is effectively the variance if the procedure is governed by Poisson statistics.) The minimum chi-square method of model fitting consists of minimizing the chi-squared statistic by varying the parameters of the model. The premise on which this technique is based is simply that the model is assumed to be qualitatively correct, and is adjusted to minimize (via χ2 ) the differences between the Ei and Oi which are deemed to be due solely to statistical fluctuations. In practice, the parameter search is easy enough as long as the number of parameters is less than four; if there are four or more, then sophisticated search procedures may be necessary. The appropriate number of degrees of freedom to associate with χ2 for k bins and N parameters is ν = k −1−N . The essential issue, having found appropriate parameters, is to estimate confidence limits (Section 3.1) for them. The answer is as given by Avni 1976; the region of confidence (significance level α) is defined by χ2α = χ2min + ∆(ν, α) where ∆ is from Table 6.1. (It is interesting to note that (a) ∆ depends only on the number of parameters involved, and not on the goodness of fit (χ2min ) actually achieved, and (b) there is an alternative answer given by Cline & Lesser (1970) which must be in error: the result obtained by Avni has been tested with Monte Carlo experiments by Avni himself and by M. Birkinshaw (personal communication).) Table 6.1. Chi-square differences (∆) above minimum Significance α 0.68 0.90 0.99
Number of parameters 1
2
3
1.00 2.71 6.63
2.30 4.61 9.21
3.50 6.25 11.30
6.4 The minimum chi-square method
125
The model to describe an observed distribution (Fig. 6.6, left) requires two parameters, γ and k. Contours of χ2 resulting from the parameter search are shown in Fig. 6.6 (right). When the Avni prescription is applied, it gives χ20.68 = χ2min + 2.30, for the value corresponding to 1σ (significance level = 0.68); the contour χ20.68 = 6.2 defines a region of confidence in the (γ, k) plane corresponding to the 1σ level of significance. (Because the range of interest for γ was limited from other considerations to 1.9 < γ < 2.4, the parameter search was not extended to define this contour fully.)
EXAMPLE
Fig. 6.6. An example of model fitting via minimum χ2 . The object of the experiment was to estimate the surface-density count [N (S) relation; see Section 6.1, Fig. 6.1] of faint extragalactic sources at 5 GHz, assuming a power-law N (> S) = KS −(γ−1) , γ and K to be determined from the distribution of background deflections, the so-called P(D) method, Section 7.6. The histogram of measured deflections is shown left, together with the curve representing the optimum model from minimizing χ2 . Contours of χ2 in the γ − K plane are shown right, with χ2 indicated for every second contour.
There are three good features of the minimum chi-square method, and two bad and ugly ones. The good: (i) Because χ2 is additive, the results of different datasets that may fall in different bins, bin sizes, or that may apply to different aspects of the same model, may be tested all at once. (ii) The contribution to χ2 of each bin may be examined and regions of exceptionally good or bad fit delineated. (iii) One of the finest features of the method is that you get model testing for free. Table A2.6 indicates probabilities of χ2 for given degrees of freedom. It is to be hoped that the model comes out
126
Data modelling; parameter estimation with a value of order 0.50; indeed the peak of the χ2 distribution is ∼ (number of degrees of freedom) when ν ≥ 4 (Fig. 5.4). In the example above, there are seven bins, two parameters, and the appropriate number of degrees of freedom is therefore 4. The value of χ2min is about 4, just as one would have hoped, and the optimum model is thus a satisfactory fit.
The bad and downright ugly: (i) Low bin-populations in the chi-square sums will cause severe instability. As a rule of thumb, 80 per cent of the bins must have Ei > 5. As for the chi-square test, it does not work for small numbers. (ii) Finally it is important to repeat the mantra: data binning is bad. In general, it loses information and efficiency. What is worse is the bias it can cause. Just consider a skewed distribution with rather few data defining it – the consequent need for wide bins may ‘erase’ the skewness entirely. 6.5 Monte Carlo modelling 6.5.1 Monte Carlo generators By now one truth will have dawned – there are many occasions in hypothesis testing and model fitting when it is essential to have simple recourse to a set of numbers distributed perhaps how we guess the data might be. We may wish to test a test to see if it works as advertised; we might need to test efficiency of tests; we might wish to determine how many iterations we require; or we might even want to test that our code is working. We need random numbers, either uniformly distributed, or drawn randomly from a parent population of known frequency distribution. It is vital not to compromise the tests with bad random data. Numerical Recipes (Press et al. 1992) presents a number of methods, from single expressions to powerful routines. A key issue is cycle length; how long is it before the pseudo-random cycle is repeated? (Or, how many random numbers do you need?) In these respects it is very necessary to understand the characteristics of the generator. Moreover it is essential to follow the prescribed implementation precisely. It may be tempting to try some ‘extra randomizing’, for example by combining routines or by modifying seeds. Be very scared of any such process.
6.5 Monte Carlo modelling
127
Finally it is easy to forget that the routines generate pseudo-random numbers. Run them again from the same starting point and you’ll get the same set of numbers. With these points in mind for the randomnumber generator for uniform deviates over the range 0 − 1, consider the following four aspects of random-number generation. 1. How do we draw a set of random numbers following a given frequency distribution? Suppose we have a way of producing random numbers that are uniformly distributed, in say the variable α; and we have a functional form for our frequency distribution dn/dx = f (x). We need a transformation x = x(α) to distort the uniformity of α to follow f (x). But we know that dn dn dα dα = = (6.15) dx dα dx dx and as dn/dα is uniform, thus
α(x) =
x
f (x) dx,
(6.16)
from whence the required transformation x = x(α).
Thus the example in Section 6.1: the source-count random distribution is f (x) dx = −1.5x−2.5 dx, a ‘Euclidean’ differential source count. Here dα = −1.5x−2.5 dx, α = x−1.5 , and the transformation is x = f −1 (α) = α1/1.5 .
EXAMPLE
2. The very same procedure works if we do not have a functional form for f (x) dx. If this is a histogram, we need simply to calculate the integral version, and perform the reverse function operation as above.
Fig. 6.7 shows an example of choosing uniformly distributed random numbers and transforming them to follow the frequency distribution prescribed by a given histogram.
EXAMPLE
3. How do we draw numbers obeying a Gaussian distribution? The prescription above is all very well, and works when integration of the function can be done; it can’t in many cases, the Gaussian being an
128
Data modelling; parameter estimation
Fig. 6.7. An example of generating a Monte Carlo distribution following a known histogram. Left: the step-ladder histogram, with points from 2000 trials, produced by (a) integrating the function (middle) and (b) transforming the axes to produce f −1 of the integrated distribution (right). The points with √ N error bars in the left diagram are from drawing 2000 uniformly distributed random numbers and transforming them according to the right diagram.
obvious one. Of course we could evaluate the integral for example by Monte Carlo methods as described below, but computationally this is ridiculous should we want a large number of deviates. There is thus another method, the rejection method, of generating random numbers to a prescription starting with uniform deviates. The method is computationally expensive relative to the integral transform method; but for something like a Gaussian, not prohibitively so; and it can be coded in just a few lines. Details are described in Lyons (1986) and Press et al. (1992). 4. How do we generate numbers obeying a bivariate (or even multivariate) Gaussian, with given σi and ρi ? This is crucial for testing many tests or model-fitting routines (or for generating Fig. 6.2); and thanks
6.5 Monte Carlo modelling
129
to our discussion of error matrices in Section 4.2 and PCA in Section 4.5, quite simple to formulate: • Set up the covariance matrix. (For the bivariate case, the error matrix is e1,1 = σx2 , e2,1 = e1,2 = cov[x, y] = ρσx σy , e2,2 = σy2 , as we have seen.) • Find the eigenvalues and eigenvectors of the covariance matrix. • Combine the eigenvectors, the column vectors, into the transformation matrix T , the matrix that diagonalizes the covariance matrix. • Then draw (x , y ) Gaussian pairs, uncorrelated, with variances equal to the two eigenvalues. Compute the (x, y) pairs according to x x = [T ] . (6.17) y y The points in Fig. 6.2 were obtained in this manner.
6.5.2 Monte Carlo integration One very important use of Monte Carlo is integration. This is a technical subject, well covered in Evans & Swartz (1995) and Chib & Greenberg (1995). A more technical reference is O’Ruanaidh & Fitzgerald (1996). Many-dimensional numerical integration is a big problem for Bayesian methods and so we will introduce some terminology and ideas here very briefly. Suppose we have a probability distribution f (x) defined for a ≤ x ≤ b. If we draw N random numbers X, uniformly distributed between a and b, then we have
b 1 f (x) dx f (Xi ). (6.18) N i a This is Monte Carlo integration. If the Xi are drawn from the distribution f itself, then obviously they will sample the regions where f is large and the integration will be more accurate. This technique is called importance sampling. So, in a Bayesian context, we would like to be able to generate random numbers from a probability distribution f /C where C is an unknown normalizing factor. Further, f will in general be a multivariate distribution (if it wasn’t, we could use deterministic numerical integration). The workhorse method for obtaining random numbers in this situation is the Metropolis algorithm or its cousin, the Metropolis–Hastings
130
Data modelling; parameter estimation
algorithm. This is a very simple method, which copies the way in which physical systems, in thermal equilibrium, will populate their distribution function. It produces a string of related random numbers called a Markov chain. The enormous advantage of the method is that it works when we do not know the normalization. Indeed, we nearly always want to find the normalization. The simplest implementation of the Metropolis algorithm is onedimensional. What if we want random numbers from a multivariate f (α1 , α2 , γ, . . .)? This is a much more likely application in a Bayesian context. Here we use the Gibbs sampler. This is actually one version of a multidimensional Metropolis algorithm (Chib & Greenberg 1995). We guess a starting vector (α0 , β0 , γ0 , . . .) and then draw α1 from f (α0 , β0 , γ0 , . . .). Next we draw β1 from (α1 , β0 , γ0 , . . .) and then γ1 from (α1 , β1 , γ0 , . . .); and so on. After we have cycled through all the variables once, we have our first multivariate sample. Obviously the first sample will be strongly influenced by the initial guess, and a number of iterations are necessary before burn-in is complete and the procedure is in a stationary state. The same applies to the Metropolis algorithm, which starts from a ‘seed’ value. The combination of the Metropolis algorithm and the Gibbs sampler equips us to perform the multidimensional integrations we often need in Bayesian problems. You should be aware that there is considerable technical debate around the question of how long burn-in will last in particular cases. If you want to use Monte Carlo Markov chain integration, check the references and make sure you have tested your random numbers in all the standard ways.
6.6 Bootstrap and jackknife In some data-modelling procedures, confidence intervals for the parameters fall out of the procedure. But are these realistic? And what about the procedures where they do not? Computer power can provide the answer, with the bootstrap method invented by Efron (1979); see also Diaconis & Efron (1983) and Davison & Hinkley (1997). It apparently gives something for nothing, and Efron so named it from the image of lifting oneself up by one’s own bootstraps. The method is so blatant (described, for example, in Numerical Recipes as ‘quick-and-dirty Monte Carlo’) that it took some time to gain
6.6 Bootstrap and jackknife
131
respectability, but the foundations are now secure (see, e.g. LePage & Billiard 1993; Efron & Tibshirani 1993). Suppose the sample consists of N data points, each consisting of one or more numbers (e.g. single measurements, or x, y pairs), and we wish to ascertain the error on a parameter estimated from these data points (e.g. mean, or slope of a best-fit). We calculate the parameter using a modelling process such as one of those described above. We then ‘bootstrap’ to find its uncertainty, as follows: (i) Label each data point; (ii) Draw at random a sample of N with replacement (simply done by computer with a random-number generator); (iii) Recalculate the parameter. (iv) Repeat this process as many times as possible. That’s it. Provided that the data points are independent (in distribution and in order), the distribution of these recalculated parameters maps the uncertainty in the estimate from the original sample.
Bhavsar (1990) described how ideally suited the bootstrap is to estimating uncertainty in measuring the slope of the angular twopoint correlation function for galaxies. This function w(θ) (Section 9.4) measures the excess surface density over that expected from a uniform independent and random distribution at angular scales θ. The data points are the (x, y) pairs of galaxy coordinates on the sky, and the difficulty in estimating the accuracy of this slope is even more notorious than that of estimating the slope of the counts of radio sources. The reason is similar: √ N error bars are readily assigned, but they are not independent; and unlike the case of source counts for which a differential version is possible, there is no ready way of assessing the significance of the correlated errors in a correlation function. Figure 6.8 shows an example of such a two-point correlation function estimate, part of a search for clustering in the distribution of radio sources on the sky (Wall, Rixon & Benn 1993).
EXAMPLE
The bootstrap is ideal for computing errors in a PCA analysis. It is a good way of telling you if any of the principal components has been detected above the sampling error.
132
Data modelling; parameter estimation
Fig. 6.8. A bootstrap application. (a) The two-point correlation function for 2812 radio sources with extended radio structure, from the White–Becker catalogue of the NRAO 1.4-GHz survey of the northern sky. A least-squares fit gives a slope of −0.19. (b) The distribution of slopes obtained in bootstraptesting the sample with 1000 trials. The mean slope is −0.157, while the rms scatter is ±0.082; the slope is less than zero (i.e. signal is present) for 96.8 per cent of the trials.
The bootstrap takes us back to the quotation starting this chapter. If errors are not well known, it is still possible to ascertain errors on a model. Moreover the errors may be known well; but as in the above example, their significance in terms of defining a model may not be understood. In either case it is possible to bootstrap one’s way to safety. The jackknife is a rather similar technique to the bootstrap, but much older, first described by Tukey (one of the inventors of the FFT) in 1958. The algorithm is again quite simple. Suppose we are interested in some function f (X1 , X2 , . . .) which depends on the N observations Xi . Usually this will be because f is a useful estimator of a parameter α. Thus we have α ˆ = f (X1 , X2 , . . .). The jth partial estimate is obtained by deleting the jth element of the dataset: α ˆ j = f (X1 , X2 , . . . , Xj−1 , Xj+1 , . . . , XN ), giving N partial estimates. The next step (and the crucial one) is to define the pseudo-values α ˆ j∗ = N α ˆ − (N − 1)ˆ αj , and finally the jackknifed estimate of α is the simple average of the
6.7 Models of models, and the combination of datasets
133
pseudo-values α ˆ∗ =
N 1 ∗ α ˆ . N i=1 j
(6.19)
The great merit of the jackknife is that it removes bias. Often the bias will depend inversely on the sample size (a simple example of this is the maximum-likelihood estimate for the variance of a normal distribution) and the jackknifed estimate will not contain this bias. In general, we can construct an mth-order jackknifed estimate by removing m observations at a time, and this will eliminate bias that depends on 1/N m . For estimators which are asymptotically Normal (e.g. maximumlikelihood estimators) it is useful to calculate the sample variance on the pseudo-values, which is 1 (σ ∗ )2 = (ˆ αj∗ − α ˆ ∗ )2 . (6.20) N (N − 1) j This can be used to give a confidence interval on α − α∗ which is distributed according to σ ∗ t with t having N − 1 degrees of freedom. This works to the extent that Normality has been obtained. In practice it is easier to use a bootstrap for confidence intervals, because the assumption of Normality is not needed. If the jackknife intervals can be checked with a bootstrap, they are of course much less computationally intensive to calculate. 6.7 Models of models, and the combination of datasets Having the correct model is essential, as otherwise both deduced parameters, and errors on them, will be wrong. Frequently, however, we are in a circular type of reasoning where we guess the model and then try to assess if the deduced parameters are reasonable. A useful way of expanding the set of models, as an insurance policy against having the wrong one, is to use hierarchical models. These in turn make use of the even more impressively named hyperparameters. It turns out that, in addition to helping with modelling, these notions are useful in the familiar problem of combining sets of data which have different levels of error. The idea of the hierarchical model can be illustrated by our earlier example, where we needed to include some kind of offset in the model for each of our flux measurements. Each term in the likelihood function
134
Data modelling; parameter estimation
took the form
2 − Si − β + kfi−γ 1 √ . exp 2 2π kfi−γ 2 kfi−γ
We are assuming that the offset error β is the same for each measurement. Before, we supposed that the distribution of β was normal, with a known mean and standard deviation – quite a strong assumption. Suppose we knew only the standard deviation, but the mean µ was unknown. The likelihood is then 2 − Si − β + kfi−γ −(β − µ)2 1 √ exp exp 2 2σβ2 2π kfi−γ 2 kf −γ i i
where µ is now a hyperparameter, described (appropriately enough) by a hyperprior. So, for hierarchical models, Bayes’ theorem takes the form prob(α, θ | Xi ) ∝ L(Xi | α)prob(α | θ)prob(θ)
(6.21)
where as usual Xi are the data and θ is the hyperparameter (and may of course be a vector). If we integrate out θ, we get a posterior distribution for the parameter α which includes the effect of a range of models.
In our radio spectrum example, we make a simple hierarchical model as described above. Take the standard deviation σβ =
and the prior prob(µ) = constant. We compute the likelihood surface by marginalizing over both µ and β; these integrations are not too bad because we have Gaussians, and because we integrate from −∞ to ∞. (More realistic integrations, over finite ranges, get very messy.) In Fig. 6.9 we see the likelihood surface for K and γ, compared to the previous ‘strong’ model for which we knew µ. There is a tendency, not unexpected, for flatter power laws to be acceptable if we do not know much about µ.
EXAMPLE
In a more elaborate form of a hierarchical model, we can connect each datum to a separate model, with the models being joined by an overarching structural relationship. In symbols, Bayes then reads prob(αi , θ | Xi ) ∝ L(Xi | αi )prob(αi | θ)prob(θ).
(6.22)
In a common type of model we may have observations Xi drawn from Gaussians of mean µi , with a structural relationship that tells us that the
6.7 Models of models, and the combination of datasets
135
1.4
K
1.2 1 0.8 0.6 0.4
0.6
0.8
1
1.2
1.4
Fig. 6.9. The log likelihoods for the two models; the black contours are for the hierarchical model and the dashed contours are for known µ.
µi are in turn drawn from a Gaussian of mean, say, θ. This is a weaker model than the first sort we considered, because we have allowed many more parameters, linked only by a stochastic relationship. In the case of Gaussians there is quite an industry devoted to this type of model; see Lee (1997) for details.
Back to our power-law spectrum. If we allow a separate offset βi at each frequency, then each term in the likelihood product takes the form
EXAMPLE
−(βi − µ)2 exp 2σβ2
2 − Si − βi + kfi−γ 1 √ exp 2 2π kfi−γ 2 kfi−γ
and we take again the usual (very weak) prior prob(µ) = constant. Marginalizing out each βi by an integration is then exactly the same task for each i, and having done this we can compare the likelihood contours with the very first model of these data (no offsets allowed). The likelihood contours of Fig. 6.10 are very instructive. The hierarchical model, by allowing a range of models, has moved the solution away from the well-defined (but wrong) parameters of the no-offset model. The hierarchical likelihood in fact peaks quite close to the true values of (k, γ) but the error bounds on these parameters are much wider.
136
Data modelling; parameter estimation
Fig. 6.10. The log likelihoods for the two models; the black contours are for the simplest model, with no provision for offsets; the dashed contours are for the weak hierarchical model, allowing separate offsets at each frequency.
This is a general message; allowing uncertainty in our models may make the answers apparently less precise, but it is an insurance against well-defined but wrong answers from modelling. Broadening the range of models is a useful technique in combining data. To see this, let us revise the idea of weights. The optimum weight for an observation of standard deviation σ is just 1/σ 2 (see the exercises). This weight turns up naturally in modelling using minimum-χ2 . Suppose we have data Xi , of standard deviation σx , and some other data Yi of standard deviation σy . Then, to fit to some model function µ(α1 , α2 , . . .) we minimize 2
χ =
N (Xi − µ)2 i=1
σx2
+
M (Yi − µ)2 i=1
σy2
and it is obvious how the different datasets are weighted. Quite often the quoted error levels on data are wrong; it is no small task to make accurate error estimates. One simple way of dealing with this is simply to tinker with the σs in the χ2 so that the minimum value comes out to be about N + M . This can be a useful technique but of course it is rather arbitrary how we allocate the tinkering between σx and σy .
6.7 Models of models, and the combination of datasets
137
Let us broaden our model by allocating weights ξx and ξy to these datasets. This is a hierarchical model, and the weights are hyperparameters (Hobson, Bridle & Lahav 2002). On the assumption of Gaussian residuals, the likelihood function is then L(Xi , Yi | α, β, . . . , ξx , ξy ) ∝
1 N/2 M/2 ξx σxN ξy σyM
(Xi − µ)2 × exp − ξx 2σx2 i=1 M (Yi − µ)2 × exp − . ξy 2σy2 i=1
N
(6.23)
Bayes’ theorem will now tell us the posterior probability distribution for the parameters of our model µ, plus the weights. It would be nice to marginalize out the weights, as in this context they are nuisance parameters. The tidy aspect of this approach is that it is one of the rare cases in which we have a convincing (uncontroversial?) prior to hand. Hobson, Bridle & Lahav (2002) show that, on the assumption that the mean value of the weight is unity (perhaps an idealistic assumption), we have simply prob(ξ) = exp(−ξ).
(6.24)
This is derived by the method of maximum entropy, as described in, for example, Jaynes (2003). Carrying out the integration over the ξ’s is easy, and we find the posterior probability for our problem to be prob(α1 , α2 , . . . | Xi , Yi ) ∝
1 1 σxN σyM 1
× 2+
N
(Xi −µ)2 2 i=1 2σx
1
× 2+
M
(Yi −µ)2 i=1 2σy2
×prob(α1 , α2 , . . .).
N/2+1
M/2+1 (6.25)
138
Data modelling; parameter estimation
Here (Fig. 6.11) are two noisy spectra of a single line. Both are alleged to have the same noise level, σ = 5, but one is slightly worse and is not centred at zero, unlike the better one. For simplicity, let us assume that we know the line to be Gaussian and only its position is unknown. Combining the data, taking the quoted errors at face value, we get a log likelihood for the line centre which peaks some way away from zero. If our prior on the line centre is diffuse, the posterior probability is proportional to the likelihood. Including the data weights as hyperparameters, we get a simple answer after marginalization, shown in Fig. 6.12; the posterior probability for the line centre shows two clear peaks, the larger at zero (the good data) and the lesser at 2 units (the poorer data).
EXAMPLE
Fig. 6.11. The two synthetic spectra which are our input data.
− 200
Posterior
Log likelihood
0
− 400 − 600 − 800 −2
−1
0 1 Line centre
2
3
3.5 3 2.5 2 1.5 1 0.5 0
−1
0 1 Line centre
2
3
Fig. 6.12. The log likelihood function for the combined, unweighted data (left) and the posterior distribution for the line centre, after marginalizing out the weights (right).
Since the weights are an amplification of our model, we may want to know if they ought to be included; this can be calculated in the usual
Exercises
139
way by computing the odds in favour of or against the more complex model. To do this we need to keep track of all the constants we have elided so far. Here is the full set of equations, for a multivariate Gaussian model for the data. Let us index each (homogeneous) set of Ni data by i, and call the i and the model vector µ covariance matrix Ci , the data vector X i . µi depends on the parameters of interest. Abbreviating i − µi ) χ2i = (Xi − µi )T Ci−1 (X the multivariate Gaussian model for the ith dataset is, as usual, i | µi , no weights) = prob(X
1 (2π)Ni /2
| Ci |1/2
exp(− 12 χ2i ) prob(µi ).
Introducing a weight simply means multiplying the covariance matrix by a factor ξi . The multivariate model for the ith dataset, after marginalizing over the weight parameter with respect to the exponential prior, is just Ni Ni /2+1 2Γ + 1 1 2 i | µi , weights) = prob(X prob(µi ). π Ni /2 | Ci |1/2 2 + χ2i (6.26) Each of these distributions depends on the parameters of the model. The odds in favour of weighting the data entail integrating over the parameters (let us abbreviate this by α ), taking account of any priors prob(α), and then forming the ratio
i | µi , weights) prob(α)prob(X . i | µi , no weights) prob(α)prob(X
α α
(6.27)
Exercises 6.1
6.2
Covariance matrix. Consider N data Xi , drawn from a Gaussian of mean µ and standard deviation σ. Use maximum likelihood to find estimators of both µ and σ, and find the covariance matrix of these estimates. Weighting data. Show that the optimum weight for an observation of standard deviation σ is just 1/σ 2 . This weight turns up naturally in modelling using minimum-χ2 .
140 6.3
6.4
6.5
6.6
6.7
Data modelling; parameter estimation MLE and power laws. In the example in Section 6.1 we fit a power law truncated at the faint end, and assume we know where to cut it off. What happens if you try to infer the faint-end cutoff by ML as well? Formulate this problem at least. Univariate random numbers. Work out the inverses of the integral functions required to generate (a) f (x) = 2x3 , (b) a power law, representative of luminosity functions, f (x) = x−γ . Use these results to produce random experiments following these probabilities by drawing 1000 random samples uniformly distributed between 0 and 1; verify by comparison with the given functions. Multivariate random numbers. (a) Give the justification for why the prescription (Section 6.5) for generating (x, y) pairs following a bivariate Gaussian of given variances and correlation coefficient is correct. (b) Using a Gaussian Monte Carlo generator, find 1000 (x, y) pairs following a given prescription, i.e. σx2 , σy2 and ρ. Plot these on contours of the bivariate probability distribution, as in Fig. 6.2, to check roughly that the prescription works. (c) Find the error matrix for the (x, y) pairs to verify that the prescription works. Monte Carlo integration. The Gaussian or Normal distribution function 1 x2 √ exp − 2 2σ σ 2π does not have an analytic integral form. Use Monte Carlo integration to find erf, the so-called error function of Table A2.1. Show that (a) approximately 68 per cent of its area lies between ±σ, and (b) that the total area under the curve is unity. Maximum likelihood estimates. Find an estimator of µ when the distribution is (a) prob(x) = exp(−|x − µ|) and (b) the Poisson prob(n) = µn
6.8
e−µ . n!
Least squares linear fits. Derive the ‘minimum distance’ OLS for errors in both x and y, assuming Gaussian errors.
Exercises 6.9
6.10
141
Marginalization. Using the data supplied, use maximum likelihood to find the distribution of the parameters of a fitted Gaussian plus a baseline. Test to see how the estimates are affected by marginalizing out the baseline parameters. The jackknife. Using the MLE for a power-law index (Section 6.1), work out and compare the confidence intervals with the analytic result from that section using the jackknife and bootstrap tests. Check how the results depend on sample size.
7 Detection and surveys
Watson, you are coming along wonderfully. You have really done very well indeed. It is true that you have missed everything of importance, but you have hit upon the method. . . (Sherlock Holmes in ‘A Case of Identity’, Sir Arthur Conan Doyle)
‘Detection’ is one of the commonest words in the practising astronomers’ vocabulary. It is the preliminary to much else that happens in astronomy, whether it means locating a spectral line, a faint star or a gamma-ray burst. Indeed of its wide range of meanings, here we take the location, and confident measurement, of some sort of feature in a fixed region of an image or spectrum. When a detection is obvious to even the most sceptical referee, statistical questions usually do not arise in the first instance. The parameters that result from such a detection have a signal-to-noise ratio so high that the detection finds its way into the literature as fact. However, elusive objects or features at the limit of detectability tend to become the focus of interest in any branch of astronomy. Then, the notion of detection (and non-detection) requires careful examination and definition. Non-detections are especially important because they define how representative any catalogue of objects may be. This set of non-detections can represent vital information in deducing the properties of a population of objects; if something is never detected, that too is a fact, and can be exploited statistically. Every observation potentially contains information. If we are resurveying a catalogue at some new wavelength, each observation constrains the energy from the object to some level. Likewise, surveying unmapped regions of sky yields information even when there are apparently no detections. In both cases population properties 142
7.1 Detection
143
can be extracted, even though individual objects remain obscured in the fog of low signal-to-noise ratio. This chapter will examine detection, first in the context of the use to which we will put detected objects; it moves on to consider the usefulness of non-detections in deducing properties of populations; and finally it examines notions of detection which say little about individual objects, but which focus instead on population-level properties. In many experiments, we wish to define wide distributions of widely spread parameters: the initial mass function, luminosity function, and so on. We may approach these from the point of view of ‘detections’ and ‘non-detections’ (the catalogue point of view) or we may attempt to extract the distributions directly from the data, without the notion of detection ever intruding.
7.1 Detection Detection is a model-fitting process. When we say ‘We’ve got a detection’ we generally mean ‘We have found what we were looking for’. This is obvious enough at reasonable signal-to-noise. In examining a digital image, for example, detection of stars (point-like objects) is achieved by comparing model point-spread functions to the data. In the case of extended objects, a wider range of models is required to capture the possibilities. In all cases, a clear statistical model is required. The noise level (or expected residuals √ from the model) may be expected in many cases to follow Poisson ( N ) statistics, or, for large N , Gaussian statistics. The statistics depend on more than the physical and instrumental model. How were the data selected for fitting in the first place? We will see for example that picking out the brightest spot in a spectrum (Section 8.6.1) means that we have a special set of data. The peak pixel, in this case, will follow the distribution appropriate to the maximum value of a set of, say, Gaussian variables. Adjacent pixels will follow an altogether less well-defined distribution; Monte Carlo simulation may be the only way forward. Indeed much evaluation of detection is done with simulation. ‘Model sources’ are strewn on the image or spectrum, and the reduction software is given the job of telling us what fraction is detected. These essential large-scale techniques are very necessary for handling the detail of how the observation was made. Evaluating detection level in radio-astronomy
144
Detection and surveys
synthesis images is an example. The noise level at any point depends at least on gains of all antennas, noise of each receiver, sidelobes from whatever sources happen to be in the field of view, map size, weighting and tapering parameters, the ionosphere, cloud, and so on. Modelling all this is not just impossible from a computational point of view – vital input data simply are not known. Although complex and varied issues are involved, the basic notions and algorithms of detection remain just as relevant as in apparently simpler cases. The basic problem from a statistical point of view is the problem of modelling, as discussed in the last chapter. A full Bayesian approach is desirable but computationally intensive and certainly not practical in a surveying application. An entry point to the Bayesian literature on this subject is Hobson & McLachlan (2003). We may need a simpler method, and a classical approach is useful. Firstly, we have to ask: what do we really want from the survey we are planning? Are we more concerned with detecting as much as possible (completeness) or are we more worried about false detections (reliability)? Moreover, we need to know what we want to do with the ‘detections’ once we have them. Perhaps we should publish, in a catalogue, the complete set of posterior probabilities, at each location, of the observed parameters? Or just the covariance matrix, as an approximation? Or perhaps the marginalized signal-to-noise ratio, integrating away all nuisance parameters? Scientific judgement must be used to answer these questions. The more information we catalogue, the better; and in the Internet age, this is so inexpensive as to be almost mandatory. From the classical point of view, if we are trying to measure a parameter α then the likelihood sums up what we have achieved: L = prob(data | α). To be specific, suppose that α is a flux density and we wish to set a flux limit for a survey. We are only going to catalogue detections when our data exceed this limit slim . (Other quantities of astrophysical interest may need a somewhat different formulation, but the essential points remain the same.) Two properties of the survey are useful to know. (i) The false-alarm rate is the chance that pure noise will produce data above the flux limit: F(data, slim ) = prob(data > slim | α = 0).
(7.1)
The reliability is 1−F, i.e. F = 5/100 gives 95 per cent reliability. That may sound good, but note that it is the infamous 2σ result.
7.1 Detection
145
(ii) The completeness is the chance that a measurement of a real source will be above the flux limit: C(data, slim , S) = prob(data > slim | α = s).
(7.2)
These notions go back as least as far as Dixon & Kraus (1968); an interesting recent treatment is by Saha (1995). We would like to set the flux limit to maximize the completeness, and minimize the false-alarm rate. But higher completeness (or even complete completeness, slim = 0!) comes at the price of an increasing number of false detections. Moreover this definition of completeness only takes account of statistical effects. There may be other reasons for missing objects, poor recognition algorithms in particular.
Suppose our measurement is of a flux density s and the noise on the measurement is Gaussian, of unit standard deviation. The source we are observing has a ‘true’ flux density of s0 , measured in units of the standard deviation. We then have 1 (s − s0 )2 prob(s | s0 ) = √ exp − 2 2π
EXAMPLE
for the probability density of the data, given the source; and 1 −s2 prob(s | s0 = 0) = √ exp − 2 2π for the probability density of the data when there is no source. Integrating these functions from 0 to slim (Table A2.1) makes it easy to plot up the completeness against the false-alarm rate, taking the flux limit as a parameter (Fig. 7.1). High completeness does indeed go hand in hand with a high false-alarm rate. However it is apparent that there are quite satisfactory combinations for flux limits and source intensities of just a few standard deviations. In real life no one would believe this, mainly because of outliers not described by the Gaussians assumed. Exercise 7.4 asks for a repeat of this calculation using an exponential noise distribution.
The conditional probabilities we have encountered suggest taking a Bayesian approach. We have prob(data | a source is present, brightness s)
146
Detection and surveys
Completeness
1 0.7 0.5
0.3
0.005 0.01 0.05 0.1 False alarm rate
0.5
Fig. 7.1. Completeness versus false-alarm rate, plotted for source flux densities in terms of σnoise ranging from 1 unit (right) to 4 units (left). The flux limits are indicated by the dots, starting at zero on the right and increasing by one unit at a time. For example, a 4σ source and a 2σ flux limit give a false-alarm rate of 2 per cent and a completeness of 99 per cent with the Gaussian noise model.
and prob(data | no source is present). Take the prior probability that a source, intensity s, is present in the measured area to be N (s), where N (s) is a normalized distribution. This is the probability that a single source will have a flux density s. The prior probability of no source is (1 − )δ(s); δ is a Dirac delta function. Then the posterior probability density prob(a source is present, brightness s | data) is given by prob(data | s)N (s) . prob(data | s)N (s) ds + (1 − ) prob(data | s = 0) Integrating this expression over s gives the probability that a source is present, for given data.
Pursuing the previous example, take the noise distribution to be Gaussian and take the prior N (s) to be a simple uniform distribution from zero to some large flux density – a very uninformative prior! The value of reflects our initial confidence that a source is present at all, and so in many cases will be small. Figure 7.2 shows that the posterior distribution of flux density s peaks at the value of the data, as expected;
EXAMPLE
7.1 Detection 1 0.8
0.3 p detection
p detected s
0.4
147
0.2 0.1
0.6 0.4 0.2
0
0 0
2
4
6
8
10
0
1
2
3 data
4
5
6
0
1
2
3 data
4
5
6
0.5
1
0.4
0.8 p detection
p detected s
s
0.3 0.2 0.1
0.6 0.4 0.2
0
2
4
6 s
8
10
Fig. 7.2. The top left panel shows the probability of a detected source of flux density s; the curves correspond to measurements of 1–4 units (as before, a unit is one noise standard deviation). A prior = 0.05 was used. On the top right these curves are integrated to give the probability of detection at any positive flux density, as a function of the data values; the curves are for = 0.5, 0.05 and 0.005. The bottom panels show the results of the calculation for the power-law prior, truncated at 0.1 unit.
the role of is to suppress our confidence of a detection in low signal-tonoise cases. Again we see that for Gaussian noise, 4σ data points mean detection with high probability. Real life is more complicated. Using a power-law prior N (s) ∝ s−5/2 gives results rather similar to the example of Fig. 2.2, which ignored the possibility that no source might be present. The rarity of bright sources in this prior now means that we need a rather better signal-to-noise to achieve the same confidence that we have a detection.
A Bayesian treatment of detection gives a direct result; from the figures in the previous example, we may read off a suitable flux limit that will give the desired probability of detection. This is affected by the prior on the flux densities, but often we will have a robust idea of what this should be from previous survey parameters such as source counts.
148
Detection and surveys
In many cases, however, the notion of detection of individual objects is poorly defined. Images or spectral lines crowd together, even overlap as we reach fainter and fainter. Within the region we measure, several different objects may contribute to the total flux. Even if only one object is present, if the source count N (s) is steep it will be more likely that the flux we measure results from a faint source plus a large upward noise excursion, rather than vice versa. In these cases we can expect only to measure population properties – parameters of the flux-density distribution N (s). If these parameters are denoted by α then a probabilistic model for the observations, when the average number of sources per measurement area is less than 1, is prob(α | data) ∝ prob(data | s)prob(s | N, α)prob(α). s
(This is an example of a hierarchical model, discussed in Section 6.7. The quantities α are really hyperparameters.) The summation in this equation will often denote a convolution between N (s) and the error distribution; given a prior on the parameters of s we can obtain a better estimate of the distribution of the flux densities of sources. If there are many sources per measurement area (and this will often be the case for faint sources) then we are in the ‘confusion-limited’ regime. Now we need to draw a distinction between N (s), the distribution of flux densities when only one source contributes, and a more complicated distribution which takes account of the possibility that several sources may add up to give s. This complicated situation is considered in Section 7.6; the details for the simpler case are left to Exercise 7.2, and they are very similar to the previous examples. In summary, detection is a modelling process; it depends on what we are looking for, and how the answer is expressed depends on what we want to do with it next. The simple idea of a detection, making a measurement of something that is really there, only applies when signalto-noise is high and individual objects can be isolated from the general distribution of properties. At low signal-to-noise, measurements can constrain population properties, with the notion of ‘detection’ disappearing.
7.2 Catalogues and selection effects Typically, a body of astronomical detections is published in a catalogue. On the basis of some clear criterion, objects will either be listed in the
7.2 Catalogues and selection effects
149
catalogue, or not. If they are not, usually we know nothing more about them; they are simply ‘below the survey limit’. Most astronomical measurements are affected by the distance to the object. In Euclidean space, a proper motion, for a fixed velocity of the star, becomes a smaller angle inversely as the distance to that star. Apparent intensity drops off as the square of the distance. Other effects may be more subtle; the ellipticity of a galaxy becomes harder to detect, depending on distance, the blurring effect of seeing, and the detailed luminosity profile of the galaxy. The common factor in all these examples is that we measure a so-called apparent quantity X and infer an intrinsic quantity by a relationship Y = f (X, R) where R is the distance to the object in question. The function f may be complicated, for observational reasons and also because it may depend on a distance involving redshift and details of space-time geometry. We take a simple and definite case (remembering that the principles will apply to the whole range of functions f ). We observe a flux density S and infer a luminosity L given by L = SR2 ; we are considering a flat-space problem. The smallest value of S we are prepared to believe is slim ; if a measurement is below this limit, the corresponding object does not appear in our catalogue. (As before, we use upper-case letters to denote measured values of the variable written in lower case.) Our objects (call them ‘galaxies’) are assumed to be drawn from a luminosity function ρ(l), the average number of objects near l per unit volume. Using only our catalogue set of measurements {L1 , L2 , . . .}, however, we will not be able to reproduce ρ at all. Instead, we will get the luminosity distribution η, where η(l) ∝ ρ(l)V (l).
(7.3)
Crucially, V (l) is the volume within which sources of intrinsic brightness l will be near enough to find their way into our catalogue. We get η(l) ∝ ρ(l)
l slim
3/2 .
(7.4)
Obviously η will be biased to higher values of luminosity than ρ. This sort of bias occurs in a multitude of cases in astronomy, and is often called Malmquist bias.
150
Detection and surveys
The luminosity function of field galaxies is well approximated by the Schechter function γ l l ρ(l) ∝ , exp − l∗ l∗
EXAMPLE
in which we take γ = 1 and l∗ = 10 for illustration. To obtain the form of the luminosity distribution in a flux-limited survey, we multiply the Schechter function by l3/2 . The differences between the luminosity function and luminosity distribution are shown in Fig. 7.3.
log
1 0.5 0.1 0.05 0.01 0.005 0.001
1
2
5 10 log luminosity
20
50
Fig. 7.3. The luminosity function ρ (steep curve) and the (flat-space) luminosity distribution are plotted for the Schechter form of the luminosity function.
Malmquist bias is a serious problem in survey astronomy. The extent of the bias depends on the shape of the luminosity function, which may not be well known. More seriously, the bias will also be present for objects whose properties correlate with something that is biased. For example, the luminosity of giant HII regions is correlated with the luminosity of the host galaxy, so that any attempt to use the HII regions as standard candles will have to consider the bias in luminosity of the hosts. Malmquist bias arises because intrinsically bright objects can be seen within proportionately much greater volumes than small ones. Because most of the volume of a sphere is at its periphery, it follows that in a flux-limited sample the bright objects will tend to be further away than the faint ones – there is an in-built distance–luminosity correlation.
7.2 Catalogues and selection effects
151
We adopt a Schechter function with γ = 1 and l∗ = 10 for the purposes of illustration. The probability of a galaxy being at distance R is proportional to R2 , in flat space. The probability of it being of brightness l is proportional to the Schechter function. The probability of a galaxy of luminosity L, located at distance R being in our sample is 1 L < slim R2 prob(in sample) = 0 otherwise.
EXAMPLE
The product of these three probability terms is the bivariate distribution prob(l, r), the probability of a galaxy of brightness l and distance r being in our sample. This distribution is shown in Fig. 7.4; there is a clear correlation between distance and luminosity. (It is this effect that produces diagrams like Fig. 4.1.) A direct check of this is to simulate a large spherical region filled with galaxies whose luminosities are drawn from a Schechter function, and then select a flux-limited sample. (The Schechter function has to be truncated at l > 0 as it otherwise cannot be normalized.) Figure 7.5 shows the effect indicated by the contours of Fig. 7.4. 5
Luminosity
4 3 2 1 0
0.2
0.4 0.6 Distance
0.8
1
Fig. 7.4. Contour plots of the bivariate prob(l, r). The contours are at logarithmic intervals; galaxies tend to bunch up against the selection line, leading to a bogus correlation between luminosity and distance.
The luminosity–distance correlation is widespread, insidious and very difficult to unravel. It means that for flux-limited samples, intrinsic properties correlate with distance; thus two unrelated intrinsic properties will
152
Detection and surveys 5
Luminosity
4 3 2 1 0
0.2
0.4 0.6 Distance
0.8
Fig. 7.5. Results of a simulation of a flux-limited survey of galaxies drawn from a Schechter function.
appear to correlate because of their mutual correlation with distance. Plotting intrinsic properties – say, X-ray and radio luminosity – against each other will be very misleading. Much further analysis is necessary to establish the reality of correlations, or (more generally) statistical dependence. Such analyses may require detailed modelling of the detection process. Take the case of measuring the ellipticity of galaxies – distant ones may well look rounder because of the effects of seeing. As more distant galaxies seem to be more luminous as well, we are on course for deducing without evidence that round galaxies are more luminous or vice versa. A detailed model will be necessary to establish the relationship between true ellipticity, measured ellipticity, and the size of the galaxy relative to the seeing disc.
We take the same simulation as before, but attribute two luminosities to each galaxy, drawn from different Schechter functions. These might be luminosities in different colour bands, for example, and by definition are statistically independent. If we construct a flux-limited survey in which a galaxy enters the final sample only if it falls above the flux limit in both bands, we see in Fig. 7.6 that a bogus but convincing correlation emerges between the two luminosities.
EXAMPLE
Finally, we should note that an effect that competes with Malmquist bias is caused by observational error. The number of objects as a function of apparent intensity N (s), the number counts or source counts, usually
7.3 Luminosity (and other) functions
153
Log luminosity 2
50 10 5 1 0.5 0.1 0.5
1
5 10 Log luminosity 1
50 100
Fig. 7.6. Results of a simulation of a flux-limited survey of galaxies, where each galaxy has two statistically independent luminosities associated with it.
rises steeply to small values of s – there are many more faint objects than bright ones. In compiling a catalogue we in effect draw samples from the number-count distribution, forget those below slim , and convert the retained fluxes to luminosities. The effect of observational error is to convolve the number counts with the noise distribution. Because of the steep rise in the number counts at the faint end the effect is to contaminate the final sample with an excess of faint objects. (An object of observed apparent flux density is much more likely to be a faint source with a positive noise excursion than a bright source with a negative excursion.) This can severely bias the deduced luminosity function towards less luminous objects. This effect does not occur if the observational error is a constant fraction of the flux density, and the source counts are close to a power law. Many types of astronomical observations suffer from the range of problems due to Malmquist bias, parameter–distance correlation and sourcecount bias. This discussion has dealt with galaxies and luminosities for illustration; plenty of other examples could have been chosen.
7.3 Luminosity (and other) functions In this section we assume that we are dealing with a catalogue of objects, of high reliability and well-understood limits. If we are interested in some intrinsic variable l (say a luminosity), then the luminosity function ρ(l) is often important. In principle we could get an approximation to ρ by measuring Li for all of the objects in some (large) volume. In practice we
154
Detection and surveys
need another way, because high luminosities are greatly over-represented in flux-limited surveys, as we have seen. One of the best methods to estimate ρ(l) is the intuitive Vmax method (Rowan-Robinson 1968; Schmidt 1968). The quantities Vmax (Li ) are the maximum volumes within which the ith object in the catalogue could lie, and still be in the catalogue. Vmax thus depends on the survey limits, the distribution of the objects in space, and the way in which detectability depends on distance. In the simplest case, a uniform distribution in space is assumed. Given the Vmax (Li ), an estimate of the luminosity function is 1 ρˆ(Bj−1 < l ≤ Bj ) = (7.5) Vmax (Li ) Bj−1